Circuit Analysis and Design 9781607854838, 9781607854845

8,646 1,504 165MB

English Pages 795

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Circuit Analysis and Design
 9781607854838, 9781607854845

Table of contents :
CAD101.pdf
ulaby_circuits_pass4_front.pdf
ulaby_circuits_pass4_ch01.pdf
ulaby_circuits_pass4_ch02.pdf
ulaby_circuits_pass4_ch03.pdf
NewPage 135 CAD101.pdf
CAD101
ulaby_circuits_pass4_ch04.pdf
ulaby_circuits_pass4_ch05.pdf
ulaby_circuits_pass4_ch06.pdf
ulaby_circuits_pass4_ch07.pdf
ulaby_circuits_pass4_ch08.pdf
ulaby_circuits_pass4_ch09.pdf
ulaby_circuits_pass4_ch10.pdf
ulaby_circuits_pass4_ch11.pdf
ulaby_circuits_pass4_ch12.pdf
ulaby_circuits_pass4_ch13.pdf
ulaby_circuits_pass4_appA.pdf
ulaby_circuits_pass4_appB.pdf
ulaby_circuits_pass4_appC.pdf
ulaby_circuits_pass4_appD.pdf
ulaby_circuits_pass4_appE.pdf
ulaby_circuits_pass4_appF.pdf
ulaby_circuits_pass4_appG.pdf
ulaby_circuits_pass4_index.pdf

Citation preview

CIRCUIT ANALYSIS AND DESIGN Fawwaz T. Ulaby, Michel M. Maharbiz, & Cynthia M. Furse

Book companion website: CAD : cad.eecs.umich.edu

CIRCUIT ANALYSIS AND DESIGN Fawwaz T. Ulaby

The University of Michigan

Michel M. Maharbiz

The University of California, Berkeley

Cynthia M. Furse

The University of Utah

Copyright  2018 Fawwaz T. Ulaby, Michel M. Maharbiz, Cynthia M. Furse This book is published by Michigan Publishing under an agreement with the authors. It is made available free of charge in electronic form to any student or instructor interested in the subject matter.

Published in the United States of America by Michigan Publishing Manufactured in the United States of America

ISBN 978-1-60785-483-8 (hardcover) ISBN 978-1-60785-484-5 (electronic)

“book” — 2015/5/4 — 6:53 — page v — #5

To an academic, writing a book is an endeavor of love. We dedicate this book to Jean, Anissa, and Katie.

“book” — 2015/5/4 — 6:53 — page viii — #8

Brief Contents

Chapter 1

Circuit Terminology

1

Chapter 2

Resistive Circuits

Chapter 3

Analysis Techniques

115

Chapter 4

Operational Amplifiers

183

Chapter 5

RC and RL First-Order Circuits

Chapter 6 Chapter 7

50

Chapter 12

Circuit Analysis by Laplace Transform

630

Chapter 13

Fourier Analysis Technique

674

Appendix A

Symbols, Quantities, and Units

727

248

Appendix B

Solving Simultaneous Equations

729

RLC Circuits

330

Appendix C

Overview of Multisim

733

ac Analysis

385

Appendix D

Mathematical Formulas

736

Chapter 8

ac Power

459

Appendix E MATLAB® and MathScript®

Chapter 9

Frequency Response of Circuits and Filters

500

Appendix F

myDAQ Quick Reference 743 Guide

Chapter 10 Three-Phase Circuits

566

Appendix G

Answers to Selected Problems

Chapter 11

601

Magnetically Coupled Circuits

Index

738

761 767

“book” — 2015/5/4 — 6:53 — page ix — #9

Contents

Preface Chapter 1 1-1 1-2 TB1 1-3 1-4 1-5 TB2 1-6

TB5 2-7

90 94

Circuit Terminology

1

Summary

100

Overview Historical Timeline Units, Dimensions, and Notation Micro- and Nanotechnology Circuit Representation Electric Charge and Current Voltage and Power Voltage: How Big Is Big? Circuit Elements Summary Problems

2 4 9 10 15 20 25 30 35 41 42

Problems

101

Chapter 3

Chapter 2

Resistive Circuits

Overview Ohm’s Law Superconductivity Kirchhoff’s Law Equivalent Circuits Resistive Sensors Wye–Delta (Y–) Transformation The Wheatstone Bridge Application Note: Linear versus Nonlinear i–υ Relationships

115 116

3-1

Linear Circuits

116

3-2

Node-Voltage Method

117

3-3

Mesh-Current Method

123

TB6

Measurement of Electrical Properties of Sea Ice By-Inspection Methods

126

133

TB7 3-6

Linear Circuits and Source Superposition Integrated Circuit Fabrication Process Thevenin ´ and Norton Equivalent Circuits

136 140

3-7

Comparison of Analysis Methods

151

3-8

Maximum Power Transfer

151

TB8 3-9

Digital and Analog Application Note: Bipolar Junction Transistor (BJT) Nodal Analysis with Multisim

154 158

Summary

164

Problems

165

3-4

50 51 51 57 60 67 70 80 84 86

Analysis Techniques

Overview

3-5

2-1 TB3 2-2 2-3 TB4 2-4 2-5 2-6

Light-Emitting Diodes (LEDs) Introducing Multisim

3-10

129

161

“book” — 2015/5/4 — 6:53 — page x — #10

Chapter 4 4-1

Operational Amplifiers

183

Chapter 6

Overview

330 331

Overview

184

Op-Amp Characteristics

184

6-1

Initial and Final Conditions

331

6-2

Introducing the Series RLC Circuit

334

TB15 6-3

Micromechanical Sensors and Actuators Series RLC Overdamped Response (α > ω0 ) Series RLC Critically Damped Response (α = ω0 ) Series RLC Underdamped Response (α < ω0 ) Summary of the Series RLC Circuit Response The Parallel RLC Circuit

337 341

RFID Tags and Antenna Design General Solution for Any Second-Order Circuit with dc Sources Neural Simulation and Recording Multisim Analysis of Circuit Response

356 359

Summary

373

Problems

374

TB9 4-2

Display Technologies Negative Feedback

190 195

4-3

Ideal Op-Amp Model

196

4-4

Inverting Amplifier

198

4-5

Inverting Summing Amplifier

200

TB10 4-6

Computer Memory Circuits Difference Amplifier

203 206

6-5

4-7

Voltage Follower/Buffer

208

6-6

4-8

Op-Amp Signal-Processing Circuits

209

4-9

Instrumentation Amplifier

214

4-10

Digital-to-Analog Converters (DAC)

216

4-11

219

TB11 4-12

The MOSFET as a Voltage-Controlled Current Source Circuit Simulation Software Application Note: Neural Probes

4-13

Multisim Analysis

230

Summary

235

Problems

236

Chapter 5

RC and RL First-Order Circuits

225 229

248

Overview

249

5-1

Nonperiodic Waveforms

250

5-2

Capacitors

258

TB12 5-3

Supercapacitors Inductors

265 269

5-4

Response of the RC Circuit

275

5-5

Response of the RL Circuit

287

TB13 5-6

Hard Disk Drives (HDD) RC Op-Amp Circuits

293 295

TB14 5-7

Capacitive Sensors Application Note: Parasitic Capacitance and Computer Processing Speed Analyzing Circuit Response with Multisim Summary

301 305

Problems

5-8

RLC Circuits

6-4

6-7 TB16 6-8 TB17 6-9

Chapter 7

ac Analysis

346 348 349 353

363 369

385

Overview

386

7-1

Sinusoidal Signals

386

7-2

Review of Complex Algebra

389

TB18 7-3

Touchscreens and Active Digitizers Phasor Domain

393 396

7-4

Phasor-Domain Analysis

400

7-5

Impedance Transformations

403

7-6

Equivalent Circuits

410

7-7

Phasor Diagrams

413

7-8

Phase-Shift Circuits

416

7-9

Phasor-Domain Analysis Techniques

420

TB19 7-10

Crystal Oscillators ac Op-Amp Circuits

423 429

7-11

Op-Amp Phase Shifter

431

7-12

Application Note: Power-Supply Circuits

432

7-13

Multisim Analysis of ac Circuits

437

313

Summary

443

314

Problems

444

310

“book” — 2015/5/4 — 6:53 — page xi — #11

Chapter 8

ac Power

459

Overview

460

8-1

Periodic Waveforms

460

8-2

Average Power

463

TB20 8-3

The Electromagnetic Spectrum Complex Power

465 467

8-4

The Power Factor

472

8-5

Maximum Power Transfer

476

TB21 8-6

Seeing without Light Measuring Power With Multisim

477 482

Summary

485

Problems

486

Chapter 9

Frequency Response of Circuits and Filters

500

10-5

Power in Balanced Three-Phase Networks

582

TB26

Inside a Power Generating Station

586

10-6

Power-Factor Compensation

588

10-7

Power Measurement in Three-Phase Circuits

591

Summary

595

Problems

596

Chapter 11

Magnetically Coupled Circuits

601

Overview

602

11-1

Magnetic Coupling

602

TB27

Magnetic Resonance Imaging (MRI)

608

11-2

Transformers

611

11-3

Energy Considerations

615

Overview

501

9-1

The Transfer Function

501

9-2

Scaling

507

11-4

Ideal Transformers

617

TB22 9-3

Noise-Cancellation Headphones Bode Plots

509 512

11-5

Three-Phase Transformers

619

Summary

622

9-4

Passive Filters

522

Problems

623

9-5

Filter Order

530

TB23 9-6

Spectral and Spatial Filtering Active Filters

533 536

9-7

Cascaded Active Filters

538

TB24

Electrical Engineering and the Audiophile Application Note: Modulation and the Superheterodyne Receiver Spectral Response with Multisim

544

9-8 9-9

Chapter 12

Circuit Analysis by Laplace Transform

630

Overview

631

12-1

Unit Impulse Function

631

547

12-2

The Laplace Transform Technique

633

550

TB28

3-D TV

637

Summary

555

12-3

Properties of the Laplace Transform

639

Problems

556

12-4

Circuit Analysis Procedure

641

12-5

Partial Fraction Expansion

644

566

TB29

Mapping the Entire World in 3-D

648

Overview

567

12-6

s-Domain Circuit Element Models

652

10-1

Balanced Three-Phase Generators

568

12-7

s-Domain Circuit Analysis

655

10-2

Source-Load Configurations

572

12-8

662

10-3

Y-Y Configuration

574

Multisim Analysis of Circuits Driven by Nontrivial Inputs

10-4

Balanced Networks

576

Summary

665

TB25

Minaturized Energy Harvesting

577

Problems

665

Chapter 10 Three-Phase Circuits

“book” — 2015/5/4 — 6:53 — page xii — #12

Chapter 13

Fourier Analysis Technique

674 675 675 677 688

13-3

Overview Fourier Series Analysis Technique Fourier Series Representation Bandwidth, Data Rate, and Communication Circuit Applications

13-4 TB31 13-5 TB32 13-6 13-7

Average Power Synthetic Biology Fourier Transform Brain-Machine Interfaces (BMI) Fourier Transform Pairs Fourier versus Laplace

693 695 697 702 704 710

13-8 13-9

Circuit Analysis with Fourier Transform Multisim: Mixed-Signal Circuits and the Sigma-Delta Modulator Summary Problems

711 713

13-1 13-2 TB30

Appendix A Symbols, Quantities, and Units

727

Appendix B Solving Simultaneous Equations

729

Appendix C

Overview of Multisim

733

Appendix D

Mathematical Formulas

736

Appendix E

MATLAB® and MathScript®

738

Appendix F

myDAQ Quick Reference 743 Guide

Appendix G

Answers to Selected Problems

690

717 718

Index

761

767

“book” — 2015/5/4 — 6:53 — page xiii — #13

List of Technology Briefs

TB1 TB2 TB3 TB4 TB5 TB6 TB7 TB8 TB9 TB10 TB11 TB12 TB13 TB14 TB15 TB16 TB17

Micro- and Nanotechnology Voltage: How Big Is Big? Superconductivity Resistive Sensors Light-Emitting Diodes (LEDs) Measurement of Electrical Properties of Sea Ice Integrated Circuit Fabrication Process Digital and Analog Display Technologies Computer Memory Circuits Circuit Simulation Software Supercapacitors Hard Disk Drives (HDD) Capacitive Sensors Micromechanical Sensors and Actuators RFID Tags and Antenna Design Neural Simulation and Recording

10 30 57 70 90 126 136 154 190 203 225 265 293 301 337 356 363

TB18 Touchscreens and Active Digitizers TB19 Crystal Oscillators TB20 The Electromagnetic Spectrum TB21 Seeing without Light TB22 Noise-Cancellation Headphones TB23 Spectral and Spatial Filtering TB24 Electrical Engineering and the Audiophile TB25 Minaturized Energy Harvesting TB26 Inside a Power Generating Station TB27 Magnetic Resonance Imaging (MRI) TB28 3-D TV TB29 Mapping the Entire World in 3-D TB30 Bandwidth, Data Rate, and Communication TB31 Synthetic Biology TB32 Brain-Machine Interfaces (BMI)

393 423 465 477 509 533 544 577 586 608 637 648 688 695 702

Preface Welcome to Circuit Analysis and Design. As the foundational course in the majority of electrical and computer engineering curricula, an electric circuits course should serve four vital objectives: (1) It should introduce the fundamental principles of circuit analysis and equip the student with the skills necessary to analyze any planar, linear circuit, including those driven by dc or ac sources, or by more complicated waveforms such as pulses and exponentials. (2) It should start the student on the journey of circuit design. (3) It should guide the student into the seemingly magical world of domain transformations—such as the Laplace and Fourier transforms, not only as circuit analysis tools, but also as mathematical languages that are “spoken” by many fields of science and engineering. (4) It should expand the student’s technical horizon by introducing him/her to some of the many allied fields of science and technology. This book aims to accomplish exactly those objectives. Among its distinctive features are: Technology Briefs: The book contains 32 Technology Briefs, each providing an overview of a topic that every electrical and computer engineering professional should become familiar with. Electronic displays, data storage media, sensors and actuators, supercapacitors, and 3-D imaging are typical of the topics shared with the reader. The Briefs are presented at a technical level intended to introduce the student to how the concepts in the chapter are applied in real-world applications and to interest the reader in pursuing the subject further on his/her own. Technology Briefs cover applications in circuits, medicine, the physical world, optics, signals and systems, and more.

Application Notes: Most chapters include a section focused on how certain devices or circuits might be used in practical applications. Examples include power supplies, CMOS inverters in computer processors, signal modulators, and several others. Multisim and MathScript: Multisim is a SPICE circuit simulator available from National Instruments (see cad.eecs.umich.edu for details). Multisim is highlighted through many end-of-chapter demonstrations. The student is strongly encouraged to take advantage of this rich resource. The Math-Script software can perform matrix inversion and many other calculations, much like the MathWorks, Inc. MATLAB® software. myDAQ: The myDAQ board does not come with this e-book, but it can be purchased directly from National Instruments. The myDAQ is a convenient, portable measurement tool that turns a PC into a basic electrical engineering lab with a DVM, analog and digital power supplies, function generator, oscilloscope, Bode plot analyzer, and diode analyzer. A written myDAQ tutorial is available in Appendix F and online video tutorials are available at http://www.ni.com/mydaq. The book contains 53 integrative end-of-chapter problems, each intended to be solved analytically, by Multisim using software simulation, and by constructing the circuit and measuring its currents and voltages using myDAQ. The three-way complementary approach is an exceedingly valuable learning experience.

“book” — 2015/5/4 — 6:53 — page xvii — #17

Acknowledgments A science or engineering textbook is the product of an integrated effort by many professionals. Invariably, the authors receive far more of the credit than they deserve, for if it were not for the creative talents of so many others, the book would never have been possible, much less a success. We are indebted to many students and colleagues, most notably the following individuals: Richard Carnes: For his meticulous typing of the manuscript, careful drafting of its figures, and overall stewardship of the project. Richard imparted the same combination of precision and passion to the manuscript as he always does when playing Chopin on the piano. Joe Steinmeyer: For testing the Multisim problems contained in the text and single-handedly developing all of the Multisim modules on the DVD-ROMs. Shortly thereafter, Joe went to MIT at which he completed a Ph.D. in electrical engineering. Professor Ed Doering: For developing a comprehensive tutorial that includes 36 circuit problems, each of which is solved analytically, with Multisim, and with myDAQ. In addition, he created instructive video tutorials on how to use a variety of computer-based instruments, including the multimeter, oscilloscope, waveform generator, and Bode analyzer. Nathan Sawicki: For developing a tutorial (Appendix F) on myDAQ and how to build circuits using it. Rose Anderson: For developing an elegant cover design and a printable InDesign version of the book. For their reviews of the overall manuscript and for offering many constructive criticisms, we are grateful to Professors

Fred Terry and Jamie Phillips of the University of Michigan, Keith Holbert of Arizona State University, Ahmad Safaai-Jazi of Virginia Polytechnic Institute and State University, Robin Strickland of the University of Arizona, and Frank Merat of Case Western Reserve University. The manuscript was also scrutinized by a highly discerning group of University of Michigan graduate students: Mike Benson, Fikadu Dagefu, Scott Rudolph, and Jane Whitcomb. Multisim sections were reviewed by Peter Ledochowitsch. Many of the 818 end-of-chapter problems were solved and checked by students from the University of Michigan and the University of California at Berkeley. They include Holly Chiang, David Hiskens, Tonmoy Monsoor, Zachary Hargeaves, James Dunn, Christopher Lo, Chris Buonocore, and Randolf Tjandra. We thank them for their contributions. We enjoyed writing this book, and we hope you enjoy learning from it. Fawwaz Ulaby, Michel Maharbiz, and Cynthia Furse

“book” — 2015/5/11 — 17:40 — page xviii — #18

Photo Credits Page 4 Page 5

Page 6

Page 7

Page 8 Page 10 Page 11 Page 14

Page 32 Page 57 Page 57 Page 58 Page 71 Page 91 Page 126 Page 127 Page 136 Page 139

c (left) Dorling Kindersley/Getty Images; (right)  Bettmann/CORBIS; Chuck Eby (left) Chuck Eby; John Jenkins, sparkmuseum.com; IEEE History Center; History San Jos´e; (right) LC-USZ62c 39702, Library of Congress; History San Jos´e;  Bettmann/ CORBIS c (left) MIT Museum;  Bettmann/ CORBIS; Emilio Segre Visual Archives/American Institute of Physics/Science Photo Library; (right) Emilio Segre Visual Archives/American Institute of Physics/Science Photo Library (left) Courtesy of Dr. Steve Reyer; Courtesy of Texas Instruments Incorporated; NASA; Digital Equipment Corporation; (right) used with permission of SRI International; Courtesy of Texas Instruments Incorporated (left) Courtesy of IBM; Courtesy of Palra Inc., US Robotics, Inc. Miguel Rodriguez Courtesy Office of Basic Energy Sciences, Office of Science, U.S. Department of Energy From “When will computer hardware match the human brain?” by Hans Moravec, Journal of Transhumanism, Vol. 1, 1998. Used with permission National Geographic Pacific Northwest National Laboratory Courtesy of Central Japan Railway Company; Courtesy General Electric Healthcare GE Healthcare Courtesy of Khalil Najafi, University of Michigan Soomi Park Geophysical Research Letters Wendy Pyper Courtesy of Veljko Milanovic Courtesy of International Business Machines Corporation

Page 203 Page 226 Page 228 Page 229 Page 265 Page 293 Page 302 Page 303 Page 339 Page 340 Page 363 Page 364 Page 365 Page 424 Page 477 Page 478 Page 480 Page 533 Page 534 Page 546 Page 587 Page 608 Page 610 Page 648 Page 649 Page 649 Page 696 Page 702 Page 714

ZeptoBars ZYPEX, Inc. CST MICROWAVE STUDIOS Courtesy of Prof. Ken Wise, University of Michigan Science, August 18, 2006, Vol. 313 (#5789). Reprinted with permission of AAAS c Steve Allen/Brand X/Corbis  Balluff STMicroelectronics (left to right) Analog Devices; Courtesy of Prof. Khalil Najafi, University of Michigan Analog Devices (left to right) Cochlear Americas and MED-EL (left to right) John Wyatt and Medtronic Todd Kuiken, Spine-health.com, and Orthomedical NIST c Reuters/CORBIS  Suljo (top to bottom) Advance Dermatology Pocono Medical Care, Inc.; NASA/SDO Agoora.co.uk (top to bottom) Wordpress.com; SunglassWarehouse.com Martin Logan Intermountain Power Project, IECACA Image from WebPath, courtesy of Edward C. Klatt MD Emilee Minalga, Robb Merrill NASA NASA ROBYN BECK/AFP/Getty Images Aaron Chevalier and Nature (Nov. 24, 2005) Deka Corp., UC Berkeley, EPFL Courtesy of Renaldi Winoto

“book” — 2015/5/4 — 6:55 — page 1 — #1

1

CHAPTER

Circuit Terminology Contents 1-1 1-2 TB1 1-3 1-4 1-5 TB2 1-6

Overview, 2 Historical Timeline, 4 Units, Dimensions, and Notation, 9 Micro- and Nanotechnology, 10 Circuit Representation, 15 Electric Charge and Current, 20 Voltage and Power, 25 Voltage: How Big Is Big? 30 Circuit Elements, 35 Summary, 41 Problems, 42

Objectives Learn to: 

Differentiate between active and passive devices; analysis and synthesis; device, circuit, and system; and dc and ac.



Point to important milestones in the history of electrical and computer engineering.



Relate electric charge to current; voltage to energy; power to current and voltage; and apply the passive sign convention.



Describe the properties of dependent and independent sources.



Describe the operation of SPST and SPDT switches.

The iPhone is a perfect example of an integrated electronic architecture composed of a large number of interconnected circuits. Learning a new language starts with the alphabet. This chapter introduces the terms and conventions used in the language of electronics.

“book” — 2015/5/4 — 6:55 — page 2 — #2

2

CHAPTER 1

CIRCUIT TERMINOLOGY

Overview Electrical engineering is an exciting field through which we interface with the world using electrical signals. In this chapter you will learn about the basis of electrical engineering—voltage and current—where they come from, what they mean, and how to measure them. The chapter provides you the nomenclature and symbols to draw and represent electric circuits.You will also learn your first circuit analysis tool, Ohm’s law, which describes the relationship between voltage, current, and resistance. In the first section of this chapter, enjoy electrical engineering’s innovative past, and in the micro-nano Technology Brief, imagine the things you could do with it in the future. As you explore this chapter and start to pick up the tools you need in your engineering career, imagine an application that particularly interests you, and how these concepts and ideas apply to that application.

Figure 1-1: Cell phone.

Cell-Phone Circuit Architecture Electronic circuits are contained in just about every gadget we use in daily living. In fact, electronic sensors, computers, and displays are at the operational heart of most major industries, from agricultural production and transportation to healthcare and entertainment. The ubiquitous cell phone (Fig. 1-1), which has become practically indispensable, is a perfect example of an integrated electronic architecture made up of a large number of interconnected circuits. It includes a two-way antenna (for transmission and reception), a diplexer (which facilitates the simultaneous transmission and reception through the antenna), a microprocessor for computing and control, and circuits with many other types of functions (Fig. 1-2). Factors such as compatibility among the various circuits and proper electrical connections between them are critically important to the overall operation and integrity of the cell phone. Usually, we approach electronic analysis and design through a hierarchical arrangement where we refer to the overall entity as a system, its subsystems as circuits, and the individual circuit elements as devices or components. Thus, we may regard the cell phone as a system (which is part of a much larger communication system); its audio-frequency amplifier, for example, as a circuit, and the resistors, integrated circuits (ICs), and other constituents of the amplifier as devices. In actuality, an IC is a fairly complex circuit in its own right, but its input/output functionality is such that usually it can be represented by a relatively simple equivalent circuit, thereby allowing us to treat it like a device. Generally, we refer to devices that do not require an external power source in order to operate as passive devices; these include resistors, capacitors,

and inductors. In contrast, an active device (such as a transistor or an IC) cannot function without a power source. This book is about electric circuits. A student once asked: “What is the difference between an electric circuit and an electronic circuit? Are they the same or different?” Strictly speaking, both refer to the flow of electric charge carried by electrons, but historically, the term “electric” preceded “electronic,” and over time the two terms have come to signify different things:  An electric circuit is one composed of passive devices, in addition to voltage and current sources, and possibly some types of switches. In contrast, the term electronic has become synonymous with transistors and other active devices.  The study of electric circuits usually precedes and sets the stage for the study of electronic circuits, and even though a course on electric circuits usually does not deal with the internal operation of an active device, it does incorporate active devices in circuit examples by representing them in terms of equivalent circuits. An electric circuit, as defined by Webster’s English Dictionary, is a “complete or partial path over which current may flow.” The path may be confined to a physical structure (such as a metal wire connecting two components), or it may be an unbounded channel carrying electrons through it. An example of the latter is when a lightning bolt strikes the ground, creating an electric current between a highly charged atmospheric cloud and the earth’s surface.

“book” — 2015/5/4 — 6:55 — page 3 — #3

3

Human Interface, Dialing, Memory Battery Power Control

Microprocessor Control

Speech, Video, Data In

Analog-to-Digital and Digital-to-Analog Converters

Transmitter

Transmit Receive

Out

Receiver Antenna

Diplexer

Figure 1-2: Basic cell-phone block diagram. Each block consists of multiple circuits that together provide the required functionality.

Electrical engineering design is about how we use and control voltages and currents to do the things we want to do. To interface with the real world, sensors are the electrical tools that convert real world inputs—like heat, sound, light, pressure, user inputs like button presses or touch screen, motion, etc.—into voltages and currents. We then manipulate these input voltages and currents using various circuits. We may amplify them if they are too small, switch them on or off, change their frequency (filter, oscillate, modulate them), or convert them into a digital signal a computer circuit can further analyze. In the end, we want to have an output voltage or current we can use to interface back to the real world—turn on a light, buzzer, alarm, motor/actuator, or control a cell phone, car airplane, robot, medical device, etc. Electrical engineers design both the input/output (I/O) systems as well as the control and actuation circuits, and often the software and algorithms as well. Electrical engineering is about “what you can do to a voltage” and how to use it to do something important in the real world. The study of electric circuits consists of two complementary tasks: analysis and synthesis (Fig. 1-3). Through analysis, we develop an understanding of “how” a given circuit works. If we think of a circuit as having an input—a stimulus—and an output—a response, the tools we use in circuit analysis allow us to mathematically relate the output response to the input stimulus, enabling us to analytically and graphically “observe” the behavior of the output as we vary the relevant parameters of the input. An example might be a specific amplifier circuit, in which case the objective of circuit analysis might be to establish how the output voltage varies as a function of the input voltage over the full operational range of the amplifier

parameters. By analyzing the operation of each circuit in a system containing multiple circuits, we can characterize the operation of the overall system. As a process, synthesis is the reverse of analysis. In engineering, we tend to use the term design as a synonym for synthesis. The design process usually starts by defining the operational specifications that a gadget or system should meet, and then we work backwards (relative to the analysis process) to develop circuits that will satisfy those specifications. In analysis, we are dealing with a single circuit with a specific set of operational characteristics. We may employ different analysis tools and techniques, but the circuit is unique, and so are its operational characteristics. That is not necessarily the case for synthesis; the design process may lead to multiple

Analysis vs. Synthesis Analysis Circuit

Functionality

Synthesis Circuit

Specs (Design)

Figure 1-3: The functionality of a circuit is discerned by applying the tools of circuit analysis. The reverse process, namely the realization of a circuit whose functionality meets a set of specifications, is called circuit synthesis or design.

“book” — 2015/5/4 — 6:55 — page 4 — #4

4

CHAPTER 1

circuit realizations—each one of which exhibits or satisfies the desired specifications. Given the complementary natures of analysis and synthesis, it stands to reason that developing proficiency with the tools of circuit analysis is a necessary prerequisite to becoming a successful design engineer. This textbook is intended to provide you with a solid foundation of the primary set of tools and mathematical techniques commonly used to analyze both direct current (dc) and alternating current (ac) circuits, as well as circuits driven by pulses and other types of waveforms. A dc circuit is one in which voltage and current sources are constant as a function of time, whereas in ac circuits, sources vary sinusoidally with time. Even though this is not a book on circuit design, design problems occasionally are introduced into the discussion as a way to illustrate how the analysis and synthesis processes complement each other. Concept Question 1-1: What are the differences between

a device, a circuit, and a system? (See

)

CIRCUIT TERMINOLOGY

demonstrated in 1945, but computers did not become available for business applications until the late 1960s and for personal use until the introduction of Apple I in 1976. Over the past 20 years, not only have computer and communication technologies expanded at a truly impressive rate (see Technology Brief 1), but more importantly, it is the seamless integration of the two technologies that has made so many business and personal applications possible. Generating a comprehensive chronology of the events and discoveries that have led to today’s technologies is beyond the scope of this book, but ignoring the subject altogether would be a disservice to both the reader and the subject of electric circuits. The abbreviated chronology presented on the next few pages represents our compromise solution.

Chronology: Major Discoveries, Inventions, and Developments in Electrical and Computer Engineering ca. 1100 BC Abacus: the earliest known calculating device.

Concept Question 1-2: What is the difference between

analysis and synthesis? (See

)

1-1 Historical Timeline We live today in the age of electronics. No field of science or technology has had as profound an influence in shaping the operational infrastructure of modern society as has the field of electronics. Our computers and communication systems are at the nexus of every major industry. Even though no single event marks the beginning of a discipline, electrical engineering became a recognized profession sometime in the late 1800s (see chronology). Alexander Graham Bell invented the telephone (1876); Thomas Edison perfected his incandescent light bulb (1880) and built an electrical distribution system in a small area in New York City; Heinrich Hertz generated radio waves (1887); and Guglielmo Marconi demonstrated radio telegraphy (1901). The next 50 years witnessed numerous developments, including radio communication, TV broadcasting, and radar for civilian and military applications—all supported by electronic circuitry that relied entirely on vacuum tubes. The invention of the transistor in 1947 and the development of the integrated circuit (IC) shortly thereafter (1958) transformed the field of electronics by setting it on an exponentially changing course towards “smaller, faster, and cheaper.” Computer engineering is a relatively young discipline. The first all-electronic computer, the ENIAC, was built and

ca. 900 BC Magnetite: According to legend, a shepherd in northern Greece, Magnus, experienced a pull on the iron nails in his sandals by the black rock he was standing on. The rock later became known as magnetite [a form of iron with permanent magnetism]. ca. 600 BC Static electricity: Greek philosopher Thales described how amber, after being rubbed with cat fur, can pick up feathers.

1600

Electric: The term was coined by William Gilbert (English) after the Greek word for amber (elektron). He observed that a compass needle points north to south, indicating the Earth acts as a bar magnet.

1614

Logarithm: developed by John Napier (Scottish).

1642

First adding machine: built by Blaise Pascal (French) using multiple dials.

“book” — 2015/5/4 — 6:55 — page 5 — #5

1-1

HISTORICAL TIMELINE

1733

Electric charge: Charles Francois ¸ du Fay (French) discovers that charges are of two forms and that like charges repel and unlike charges attract.

1745

Capacitor: Pieter van Musschenbroek (Dutch) invented the Leyden jar, the first electrical capacitor.

1800

First electric battery: developed by Alessandro Volta (Italian).

1827

Ohm’s law: formulated by Georg Simon Ohm (German), relating electric potential to current and resistance.

1827

Inductance: introduced by Joseph Henry (American), who built one of the earliest electric motors. He also assisted Samuel Morse in the development of the telegraph.

1837

Telegraph: concept patented by Samuel Morse (American), who used a code of dots and dashes to represent letters and numbers.

5 1876

Telephone: invented by Alexander Graham Bell (Scottish-American): the rotary dial became available in 1890, and by 1900, telephone systems were installed in many communities.

1879

Incandescent light bulb: demonstrated byThomas Edison (American), and in 1880, his power distribution system provided dc power to 59 customers in New York City.

1887

Radiowaves: Heinrich Hertz (German) built a system that could generate electromagnetic waves (at radio frequencies) and detect them.

Courtesy of John Jenkins (sparkmuseum.com)

1843

1888

ac motor: invented by Nikola Tesla (Croatian-American).

1893

Magnetic sound recorder: invented by Valdemar Poulsen (Danish) using steel wire as recording medium.

Computer algorithm: original concept and plan attributed to Ada Byron Lovelace (British), the daughter of poet Lord Byron. The “Ada” software language was developed in 1979 by the U.S. Department of Defense in her honor.

“book” — 2015/5/4 — 6:55 — page 6 — #6

6 1895

1896

CHAPTER 1 X-rays: discovered by Wilhelm Rontgen ¨ (German). One of his first X-ray images was of the bones in his wife’s hands. [1901 Nobel prize in physics.]

CIRCUIT TERMINOLOGY

1917

Superheterodyne and frequency modulation (FM): invented by Edwin Howard Armstrong (American), providing superior sound quality of radio transmissions over AM radio.

1920

Commercial radio broadcasting: Westinghouse Corporation established radio station KDKA in Pittsburgh, Pennsylvania.

1923

Television: invented by Vladimir Zworykin (Russian-American). In 1926, John Baird (Scottish) transmitted TV images over telephone wires from London to Glasgow. Regular TV broadcasting began in Germany (1935), England (1936), and the United States (1939).

1926

Transatlantic telephone service established between London and New York.

1930

Analog computer: developed by Vannevar Bush (American) for solving differential equations.

1935

Anti-glare glass: developed by Katharine Blodgett by transferring thin monomolecular coatings to glass.

Radio wireless transmission: patented by Guglielmo Marconi (Italian). In 1901, he demonstrated radio telegraphy across the Atlantic Ocean. [1909 Nobel prize in physics, shared with Karl Braun (German).]

1897

Cathode ray tube (CRT): invented by Karl Braun (German). [1909 Nobel prize, shared with Marconi.]

1897

Electron: discovered by Joseph John Thomson (English), who also measured its charge-to-mass ratio. [1906 Nobel prize in physics.]

1902

Amplitude modulation: invented by Reginald Fessenden (American) for telephone transmission. In 1906, he introduced AM radio broadcasting of speech and music on Christmas Eve.

1904

Diode vacuum tube: patented by John Fleming (British).

1907

Triode tube amplifier: developed by Lee De Forest (American) for wireless telegraphy, setting the stage for long-distance phone service, radio, and television.

“book” — 2015/5/4 — 6:55 — page 7 — #7

1-1

HISTORICAL TIMELINE

1935

Radar: invented by Robert Watson-Watt (Scottish).

1944

Computer compiler: One of the earliest compilers was designed by Grace Hopper for Harvard’s Mark I computer. She retired as a rear admiral in the U.S. Navy in 1986.

7 1954

First AM transistor radio: introduced by Texas Instruments.

Courtesy of Dr. Steve Reyer

1945

1947

1955

Optical fiber: demonstrated by Narinder Kapany (Indian-American) as a low-loss, light-transmission medium.

1956

FORTRAN: developed by John Backus (American), the first major programming language.

1958

Laser: concept developed by Charles Townes and Arthur Schawlow (both Americans). [Townes shared 1964 Nobel prize in physics with Aleksandr Prokhorov and Nicolay Bazov (both Soviets).] In 1960 Theodore Maiman (American) built the first working model of a laser.

1958

Modem: developed by Bell Labs.

1958

Integrated circuit (IC): Jack Kilby (American) built the first IC on germanium, and independently, Robert Noyce (American) built the first IC on silicon.

1960

Echo: The first passive communication satellite was launched and successfully reflected radio signals back to Earth. In 1962, the first communication satellite, Telstar, was placed in geosynchronous orbit.

ENIAC: The first all-electronic computer was developed by John Mauchly and J. Presper Eckert (both American).

Transistor: invented by William Shockley, Walter Brattain, and John Bardeen (all Americans) at Bell Labs. [1956 Nobel prize in physics.]

1948

Modern communication: Claude Shannon (American) published his Mathematical Theory of Communication, which formed the foundation of information theory, coding, cryptography, and other related fields.

1950

Floppy disk: invented by Yoshiro Nakama (Japanese) as a magnetic medium for storing data.

“book” — 2015/5/4 — 6:55 — page 8 — #8

8 1960

CHAPTER 1 Microcomputer: introduced by Digital Equipment Corporation as the PDP-1, which was followed with the PDP-8 in 1965.

CIRCUIT TERMINOLOGY

1969

ARPANET: established by the U.S. Department of Defense, which later evolved into the Internet.

1970

CD-ROM: patented by James Russell (American), as the first system capable of digital-to-optical recording and playback.

1971

Pocket calculator: introduced by Texas Instruments.

Courtesy of Texas Instruments

1961

Thick-film resistor: one of 28 electronic devices patented by Otis Boykin (African-American).

1962

MOSFET: invented by Steven Hofstein and Frederic Heiman (both American), which became the workhorse of computer microprocessors.

1964

IBM’s 360 mainframe: became the standard computer for major businesses.

1965

BASIC computer language: developed by John Kemeny and Thomas Kurtz (both American).

1965

Programmable digital computer: developed by Konrad Zuse (German) using binary arithmetic and electric relays.

1968

Word processor: demonstrated by Douglas Engelbart (American), followed by the mouse pointing device and the use of a Windows-like operating system.

1971

Intel 4004 four-bit microprocessor: capable of executing 60,000 operations per second.

1972

Computerized axial tomography scanner (CAT scan: developed by Godfrey Hounsfield (British) and Alan Cormack (South African– American) as a diagnostic tool. [1979 Nobel Prize in physiology or medicine.]

1976

Laser printer: introduced by IBM.

1976

Apple I: sold by Apple Computer in kit form, followed by the fully assembled Apple II in 1977, and the Macintosh in 1984.

1979

First cellular telephone network: built in Japan: • 1983 cellular phone networks started in the United States. • 1990 electronic beepers became common. • 1995 cell phones became widely available.

1980

MS-DOS computer disk operating system: introduced by Microsoft: Windows marketed in 1985.

1981

PC: introduced by IBM.

1984

Internet became operational worldwide.

1988

First transatlantic optical fiber cable: installed between the U.S. and Europe.

1988

Touchpad: invented by George Gerpheide (American).

1989

World Wide Web: invented by Tim Berners-Lee (British) by introducing a networking hypertext system.

1996

Hotmail: launched by Sabeer Bhatia (Indian-American) and Jack Smith (American) as the first webmail service.

“book” — 2015/5/4 — 6:55 — page 9 — #9

1-2 1997

UNITS, DIMENSIONS, AND NOTATION Palm Pilot: became widely available.

9

Table 1-1: Fundamental and electrical SI units. Dimension

Unit

Symbol

Fundamental: Length Mass Time Electric charge Temperature Amount of substance Luminous intensity

meter kilogram second coulomb kelvin mole candela

m kg s C K mol cd

ampere volt ohm farad henry watt hertz

A V  F H W Hz

Electrical:

2007

White LED: invented by Shuji Nakamura (Japanese) in the 1990s. It promises to replace Edison’s lightbulb in most lighting applications.

2007

iPhone: released by Apple.

2009

Cloud computing: went mainstream.

2011

Humans vs. supercomputer: IBM’s Watson supercomputer beat the top two human contestants of Jeopardy! for a $1M prize.

2011

Text messages: 8 × 1012 (8 trillion) text messages sent worldwide.

2014

Mobile subscribers: Approximately 96% of the world population is a mobile phone subscriber (7 billion people).

Concept Question 1-3: What do you consider to be the

most important electrical engineering milestone that is missing from this historical timeline? (See )

1-2 Units, Dimensions, and Notation The standard system used in today’s scientific literature to express the units of physical quantities is the International System of Units (SI), abbreviated after its French name Syst`eme Internationale. Time is a fundamental dimension, and the second is the unit by which it is expressed relative to a specific reference standard. The SI configuration is based on the seven fundamental dimensions listed in Table 1-1, and their units are called fundamental SI units. All other dimensions, such as velocity, force, current, and voltage, are regarded as secondary because their units are based on and can be expressed in terms of the seven fundamental units. For example, electric current is measured in amps, which is an abbreviation for coulombs/ second. Appendix A provides a list of the quantities used in this book, together with their symbols and units.

Current Voltage Resistance Capacitance Inductance Power Frequency

In science and engineering, a set of prefixes commonly are used to denote multiples and submultiples of units. These prefixes, ranging in value between 10−18 and 1018 , are listed in Table 1-2. An electric current of 3 × 10−6 A, for example, may be written as 3 μA. The physical quantities we discuss in this book (such as voltage and current) may be constant in time or may vary with time.

Table 1-2: Multiple and submultiple prefixes. Prefix

Symbol

Magnitude

exa peta tera giga mega kilo

E P T G M k

1018 1015 1012 109 106 103

milli micro nano pico femto atto

m μ n p f a

10−3 10−6 10−9 10−12 10−15 10−18

“book” — 2015/5/4 — 6:55 — page 10 — #10

10

TECHNOLOGY BRIEF 1: MICRO- AND NANOTECHNOLOGY

Technology Brief 1 Micro- and Nanotechnology Scale of Things Our ability as humans to shape and control the environment around us has improved steadily over time, most dramatically in the past 100 years. The degree of control is reflected in the scale (size) at which objects can be constructed, which is governed by the tools available for constructing them. This refers to the construction of both very large and very small objects. Early tools—such as flint, stone, and metal hunting gear—were on the order of tens of centimeters. Over time, we were able to build ever-smaller and ever-larger tools. The world’s largest antenna* is the radio telescope at the Arecibo observatory in Puerto Rico (Fig. TF1-1). The dish is 305 m (1000 ft) in diameter and 50 m deep and covers nearly 20 acres. It is built from nearly 40,000 perforated 1 m × 2 m aluminum plates. On the other end of the size spectrum, some of the smallest antennas today are nanocrescent antennas that are under 100 nm long. These are built by sputtering aluminum against glass beads and then removing the beads to expose crescent-shaped antennas (Fig. TF1-2). Miniaturization continues to move forward: the first hydraulic valves, for example, were a few meters in length (ca. 400 BCE); the first toilet valve was tens of * http://www.naic.edu/general/

Figure TF1-2: Nano-crescent antenna for use in the ultraviolet range (320 nm to 370 nm wavelength). (Credit: Miguel Rodriguez.)

centimeters in size (ca. 1596); and by comparison, the largest dimension in a modern microfluidic valve used in biomedical analysis-chips is less than 100 μm! The chart in Fig. TF1-3 displays examples of manmade and natural things whose dimensions fall in the range between 0.1 nm (10−10 m) and 1 cm, which encompasses both micrometer (1 μm = 10−6 m) and nanometer (1 nm = 10−9 m) ranges. Microtechnology, which refers to our ability to manipulate matter at a precision of 1 μm or better, became possible in the 1960s, ushering in an electronics revolution that led to the realization of the laptop computer and the ubiquitous cell phone. It then took another 30 years to improve the manufacturing precision down to the nanometer scale (nanotechnology), promising the development of new materials and devices with applications in electronics, medicine, energy, and construction.

Moore’s Law

Figure TF1-1: Arecibo radio telescope.

With the invention of the semiconductor transistor in 1947 and the subsequent development of the integrated circuit in 1959, it became possible to build thousands (now trillions) of electronic components onto a single substrate or chip. The 4004 microprocessor chip (ca. 1971) had 2250 transistors and could execute 60,000 instructions per second; each transistor had a “gate” on the order of 10 μm (10−5 m). In comparison, the 2006 Intel Core had 151 million transistors with each transistor gate measuring 65 nm (6.5 × 10−8 m); it could

“book” — 2015/5/4 — 6:55 — page 11 — #11

TECHNOLOGY BRIEF 1: MICRO- AND NANOTECHNOLOGY

11

The Scale of Things – Nanometers and More Things Natural

Things Manmade 10-2 m

Ant ~ 5 mm

Dust mite

Red blood cells (~7-8 μm)

The Challenge

1,000,000 nanometers = 1 millimeter (mm) MicroElectroMechanical (MEMS) devices 10 -100 μm wide

10-4 m

0.1 mm 100 μm

10-5 m

0.01 mm 10 μm

O

Infrared

Fly ash ~ 10-20 μm

Microworld

200 μm

Human hair ~ 60-120 μm wide

Head of a pin 1-2 mm

Microwave

10-3 m

1 cm 10 mm

1,000 nanometers = 1 micrometer (μm)

O O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

O

S

S

S

S

S

S

S

S

Zone plate x-ray “lens” Outer ring spacing ~35 nm

Visible

10-6 m

Pollen grain Red blood cells

P

O

~10 nm diameter ATP synthase

Ultraviolet

Nanoworld

10-7 m

10-8 m

Fabricate and combine nanoscale building blocks to make useful devices, e.g., a photosynthetic reaction center with integral semiconductor storage.

0.1 μm 100 nm

0.01 μm 10 nm

Self-assembled, Nature-inspired structure Many 10s of nm Nanotube electrode

10-9 m

1 nanometer (nm) Soft x-ray

Carbon buckyball ~1 nm diameter Carbon nanotube ~1.3 nm diameter

DNA ~2-1/2 nm diameter

Atoms of silicon spacing 0.078 nm

10-10 m

0.1 nm

Quantum corral of 48 iron atoms on copper surface positioned one at a time with an STM tip Corral diameter 14 nm

Office of Basic Energy Sciences Office of Science, U.S. DOE Version 05-26-06, pmd

FigureTF1-3: The scale of natural and man-made objects, sized from nanometers to centimeters. (Courtesy of U.S. Department of Energy.)

perform 27 billion instructions per second. The 2011 Intel Core i7 “Gulftown” processors have 1.17 billion transistors and can perform ∼ 150 billion instructions per second. In recent years, the extreme miniaturization of transistors (the smallest transistor gate in an i7 Core is ∼ 32 nanometers wide!) has led to a number of design innovations and trade-offs at the processor level, as devices begin to approach the physical limits of classic semiconductor devices. Among these, the difficulty of dissipating the heat generated by a billion transistors

has led to the emergence of multicore processors; these devices distribute the work (and heat) between more than one processor operating simultaneously on the same chip (2 processors on the same chip are called a dual core, 4 processors are called a quad core, etc.). This type of architecture requires additional components to manage computation between processors and has led to the development of new software paradigms to deal with the parallelism inherent in such devices.

“book” — 2015/5/4 — 6:55 — page 12 — #12

12

TECHNOLOGY BRIEF 1: MICRO- AND NANOTECHNOLOGY

Transistors/Chip 109 8-Core Xeon 2.3 x 109 Dual-Core Itanium 2

109

Itanium 2 Itanium

108

Pentium 4 Pentium III

107

Pentium II Pentium II 386

106

286 8086

105

6000 8008 4004

Intel CPUs 104

8000

1970

1975

1980

1985

1990

1995

2000

2005

103 2011

Figure TF1-4: Moore’s Law predicts that the number of transistors per processor doubles every two years.

Moore’s Law and Scaling In 1965, Gordon Moore, co-founder of Intel, predicted that the number of transistors in the minimum-cost processor would double every two years (initially, he had guessed they would double every year). Amazingly, this prediction has proven true of semiconductor processors for 40 years, as demonstrated by Fig. TF1-4. In order to understand Moore’s Law, we have to understand the basics of how transistors are used in computers. Computers carry all of their information (numbers, letters, sounds, etc.) in coded strings of electrical signals that are either “on” or “off.” Each “on” or “off” signal is called a bit, and 8 bits in a row are called a byte. Two bytes are a word, and (when representing numbers) they provide 16-bit precision. Four bytes give 32-bit precision. These bits can be added, subtracted, moved around, etc., by switching each bit individually on or off, so a computer processor can be thought of as a big network of (trillions of) switches. Transistors are the basic switches in computers. We will learn more about them in Chapter 3, but for now, the important thing to know is that they can act as very tiny, very fast, very low power switches. Trillions of transistors are built directly onto a single silicon wafer (read more about how in Technology Brief 7), producing very-large-scale integrated (VLSI) circuits or chips. Transistors are characterized by their feature size, which is the smallest line width that can

be drawn in that VLSI manufacturing process. Larger transistors are used for handling more current (such as in the power distribution system for the chip). Smaller transistors are used where speed and efficiency are critical. The 22 nm processes available today can make lines and features ∼22 nm in dimension. They produce transistors that are about 100 nm on a side, switched on and off over 100 billion times a second (it would take you over 2000 years to flip a light switch that many times),† and can do about 751 billion operations per watt.‡ Even smaller, 5 nm transistors are expected to become commercially viable by 2020. The VLSI design engineer uses computer-aided design (CAD) tools to design chips by combining transistors into larger subsystems (such as logic gates that add/multiply/etc.), choosing the smallest, fastest transistors that can be used for every part of the circuit. The following questions then arise: How small can we go? What is the fundamental limit to shrinking down the size of a transistor? As we ponder this, we immediately observe that we likely cannot make a transistor smaller than the diameter of one silicon or metal atom (i.e., ∼0.2 to 0.8 nm). But is there a limit prior to this? Well, as we shrink transistors down to the point that they are † http://download.intel.com/newsroom/kits/22nm/pdfs/22nm Fun Facts.pdf ‡ https://newsroom.intel.com/servlet/JiveServlet/previewBody/2834-102-

1-5130/Intel%20at%20VLSI%20Fact%20Sheet.pdf

“book” — 2015/5/4 — 6:55 — page 13 — #13

CPU power density (W/cm2)

TECHNOLOGY BRIEF 1: MICRO- AND NANOTECHNOLOGY

100

AMD Intel Power PC

Multicores

Power dissipation 10 Single cores

Surface area

Heat flux 1 1990 1994 1998 2002 2006 2010 Year

13

Light Bulb

Integrated Circuit

100 W

50 W

106 cm2 (bulb surface area)

1.5 cm2 (die area)

0.9 W/cm2

33.3 W/cm2

Figure TF1-5: (a) Plot of heat power density generated by consumer processors over time, (b) comparison of heat power generation of a light bulb with that of a typical processor.

made of just one or a few atomic layers (∼1 to 5 nm), we run into issues related to the stochastic nature of quantum physics. At these scales, the random motion of electrons between both physical space and energy levels becomes significant with respect to the size of the transistor, and we start to get spurious or random signals in the circuit.There are even more subtle problems related to the statistics of yield. If a certain piece of a transistor contained only 10 atoms, a deviation of just one atom in the device (to a 9-atom or an 11atom transistor) represents a huge change in the device properties! This would make it increasingly difficult to economically fabricate chips with hundreds of millions of transistors. Additionally, there is an interesting issue of heat generation: Like any dissipative device, each transistor gives off a small amount of heat. But when you add up the heat produced by more than 1 billion transistors, you get a very large number! Figure TF1-5 compares the power density (due to heat) produced by different processors over time. The heat generated by single core processors increased exponentially until the mid-2000s when power densities began approaching 100 W/cm2 (in comparison, a nuclear reactor produces about 200 W/cm2 !). The inability to practically dissipate that much heat led, in part, to the development of multicore processors and a leveling off of heat generation for consumer processors. None of these issues are insurmountable. Challenges simply spur creative people to come up with innovative

solutions. Many of these problems will be solved, and in the process, provide engineers (like you) with jobs and opportunities. But, more importantly, the minimum feature size of a processor is not the end goal of innovation: it is the means to it. Innovation seeks simply to make increasingly powerful computation, not smaller feature sizes. Hence, the move towards multicore processors. By sharing the workload among various processors (called distributed computing) we increase processor performance while using less energy, generating less heat, and without needing to run at warp speed. So it seems, as we approach ever-smaller features, we simply will creatively transition into new physical technologies and also new computational techniques. As Gordon Moore himself said, “It will not be like we hit a brick wall and stop.”

Scaling Trends and Nanotechnology It is an observable fact that each generation of tools enables the construction of an even newer, smaller, more powerful generation of tools. This is true not just of mechanical devices, but electronic ones as well. Today’s high-power processors could not have been designed, much less tested, without the use of previous processors that were employed to draw and simulate the next generation. Two observations can be made in this regard. First, we now have the technology to build tools

“book” — 2015/5/4 — 6:55 — page 14 — #14

14

TECHNOLOGY BRIEF 1: MICRO- AND NANOTECHNOLOGY

Figure TF1-6: Time plot of computer processing power in MIPS per $1000. (From “When will computer hardware match the human brain?” by Hans Moravec, Journal of Transhumanism, Vol. 1, 1998.)

to manipulate the environment at atomic resolution. At least one generation of micro-scale techniques (ranging from microelectromechanical systems—or MEMS— to micro-chemical devices) has been developed that, while useful in themselves, are also enabling the construction of newer, nano-scale devices. These newer devices range from 5 nm transistors to femtoliter (10−15 ) microfluidic devices that can manipulate single protein molecules. At these scales, the lines between mechanics, electronics, and chemistry begin to blur! It is to these ever-increasing interdisciplinary innovations that the term

nanotechnology rightfully belongs. Second, the rate at which these innovations are occurring seems to be increasing exponentially! (Consider Fig. TF1-6 and note that the y axis is logarithmic and the plots are very close to straight lines.) Keeping up with rapidly changing technology is one of the exciting and challenging aspects of an engineering career. Electrical engineers use the Institute of Electrical and Electronic Engineers (IEEE) to find professional publications, workshops, and conferences to provide lifelong learning opportunities to stay current and creative (see IEEE.org).

“book” — 2015/5/4 — 6:55 — page 15 — #15

1-3

CIRCUIT REPRESENTATION

15

As a general rule, we use: • A lowercase letter, such as i for current, to represent the general case: i

may or may not be time-varying

• A lowercase letter followed with (t) to emphasize time: i(t)

is a time-varying quantity

• An uppercase letter if the quantity is not timevarying; thus: I

is of constant value (dc quantity)

• A letter printed in boldface to denote that: I has a specific meaning, such as a vector, a matrix, the phasor counterpart of i(t), or the Laplace or Fourier transform of i(t)

Exercise 1-1: Convert the following quantities to scientific notation: (a) 52 mV, (b) 0.3 MV, (c) 136 nA, and (d) 0.05 Gbits/s. Answer: (a) 5.2 × 10−2 V, (b) 3 × 105 V,

(c) 1.36 × 10−7 A, and (d) 5 × 107 bits/s. (See

)

Exercise 1-2: Convert the following quantities to a prefix

format such that the number preceding the prefix is between 1 and 999: (a) 8.32 × 107 Hz, (b) 1.67 × 10−8 m, (c) 9.79 × 10−16 g, (d) 4.48 × 1013V, and (e) 762 bits/s. Answer: (a) 83.2 MHz, (b) 16.7 nm, (c) 979 ag,

(d) 44.8 TV, and (e) 762 bits/s. (See

)

Exercise 1-3: Simplify the following operations into

a single number, expressed in prefix format: (a) A = 10 μV + 2.3 mV, (b) B = 4THz − 230 GHz, (c) C = 3 mm/60 μm. Answer: (a) A = 2.31 mV, (b) B = 3.77 THz, (c)

C = 50. (See

)

1-3 Circuit Representation When we design circuits, we first think of what we want the circuit to do (its functional block diagram), then we design

circuits to do this (a circuit diagram). We then select and lay out the components in the circuit (PCB layout) and build it. Let’s consider a capacitive-touch sensor such as the touch screen on the iphone. The circuit includes a flat conducting plate, two ICs, ), and several resistors and capacitors. When one diode ( the plate is touched by a finger, the capacitance introduced by the finger causes the output voltage to rise above a preset threshold, signifying the fact that the plate has been touched. The voltage rise can then be used to trigger a follow-up circuit such as a light-emitting diode (LED). Figure 1-4 contains four parts: (a) a block diagram of a circuit designed as a capacitortouch-sensor, (b) a circuit diagram representing the circuit’s electrical configuration, (c) the circuit’s printed-circuit-board (PCB) layout, and (d) a photograph of the circuit with all of its components. The PCB layout shown in part (c) of Fig. 1-4 displays the intended locations of the circuit elements and the printed conducting lines needed to connect the elements to each other. These lines are used in lieu of wires. The diagram in part (d) is the symbolic representation of the physical circuit. In this particular representation the resistors are drawn as rectangular boxes instead of the more familiar symbol . Designing the PCB layout and the circuit’s physical architecture is an important step in the production process, but it is outside the scope of this book. Our prime interest is to help the reader understand how circuits work, and to use that understanding to design circuits to perform functions of interest. Accordingly, circuit diagrams will be regarded as true representations of the many circuits and systems we discuss in this and the following chapters.

1-3.1

Circuit Elements

Table 1-3 provides a partial list of the symbols used in this book to represent circuit elements in circuit diagrams. By way of an example, the diagram in Fig. 1-5 contains the following elements: • A 12 V ac source, denoted by the symbol

~+− . An ac source

varies sinusoidally with time (such as a 60 Hz wall outlet). +

• A 6 V dc source, denoted by the symbol _ is constant in time (such as a battery). • Six resistors, all denoted by the symbol • One capacitor, denoted by the symbol • One inductor, denoted by the symbol

. A dc source

“book” — 2015/5/4 — 6:55 — page 16 — #16

16

CHAPTER 1

CIRCUIT TERMINOLOGY

LED Sensor

Sensor

Volt: 0

Volt: 1.5 V

(a) Block diagram

(b) Circuit diagram 5 V power supply to be connected here Metal plate Capacitor IC Diode Resistor

Output voltage (c) Printed circuit board (PCB)

(d) Actual circuit

Figure 1-4: (a) Block diagram, (b) circuit diagram, (c) printed-circuit-board (PCB) layout, (d) photograph of a touch-sensor circuit.

• An important integrated circuit known as an operational amplifier (or op amp for short), denoted by a triangular symbol (the internal circuit of the op amp is not shown).

1-3.2

Circuit Architecture

The vocabulary commonly used to describe the architecture of an electric circuit includes a number of important terms. Short, but precise, definitions follow. • Node: electrical conductor(s) or wires that connect two or more circuit elements. The node is not just a point, but includes the entire set of wires between two or more circuit elements. Nodes are color-coded in Fig. 1-5. For example, node N1 is red, N2 is green, and N3 is orange. The dot at N1

is typically used to emphasize that the wires are actually connected together. All conductors in a node always have the same voltage. • Ordinary node: an electrical connection point that connects only two elements, such as all the yellow nodes in Fig. 1-5. • Extraordinary node: node connected to three or more elements. Figure 1-5 contains four extraordinary nodes, denoted N1 through N4 , of which N4 has been selected as a reference voltage node, often referred to as the ground node. When two points with no element between them are connected by a conducting wire, they are regarded as the

“book” — 2015/5/4 — 6:55 — page 17 — #17

1-3

CIRCUIT REPRESENTATION

Ordinary node

17

υ1 = 12 cos (377t) V + ac source −

~

+ _

Op amp N3 Conducting wire

R2 Inductor

+ υ2 = 6 V dc source _

R4

Extraordinary Capacitor Branch containing R1 node C R1 R3 N1 N2

R5

R6

L Loop 1 N4

Loop 2

N4

Ground Same node Figure 1-5: Diagram representing a circuit that contains dc and ac sources, passive elements (six resistors, one capacitor, and one inductor), and one active element (operational amplifier). Ordinary nodes are in yellow, extraordinary nodes in other colors, and the ground node in black.

same node. Hence, all of the black wires together located at the bottom of the circuit in Fig. 1-5 make up node N4 . • Branch: the trace between two consecutive nodes containing one and only one element between them.

I

+ V1 _

+ V2 _

Battery

+

• Path: any continuous sequence of branches, provided that no one node is encountered more than once. The path between nodes N1 and N2 consists of two branches, one containing R3 and another containing C.

_

(a) Series circuit I1

• Loop: a closed path in which the start and end node is one and the same. Figure 1-5 contains several loops, of which two are shown explicitly.

+ V _ I2

• Mesh: a loop that encloses no other loop. In Fig. 1-5, Loop 1 is a mesh, but Loop 2 is not.

+ V _

• In series: path in which elements share the same current. As you move along a series path you encounter only ordinary nodes. Elements on these paths are in series. In Fig. 1-6(a), the two light bulbs are in series because the same current flows through both of them. Also, in Fig. 1-5, the two sources and R1 are all in series, as are R2 and L, and R3 and C.

+

Battery

_

(b) Parallel circuit Figure 1-6: Two light bulbs connected (a) in series and (b) in parallel.

“book” — 2015/5/4 — 6:55 — page 18 — #18

18

CHAPTER 1

Table 1-3: Symbols for common circuit elements. A

Table 1-4: Circuit terminology. Node: An electrical connection between two or more elements.

A

or

CIRCUIT TERMINOLOGY

Conductor Two conductors Two conductors (wire) electrically joined not joined at node A electrically

Ordinary node: An electrical connection node that connects to only two elements. Extraordinary node: An electrical connection node that connects to three or more elements. Branch: Trace between two consecutive nodes with only one element between them.

Fixed-value resistor

Variable resistor

10 V Inductor

+ _

~+−

12 V

10 V dc battery

Loop: Closed path with the same start and end node. Independent loop: Loop containing one or more branches not contained in any other independent loop.

12 V ac source

Mesh: Loop that encloses no other loops. In series: Elements that share the same current. They have only ordinary nodes between them.

+ _

6 A current source

Switch

Volts VΩ Transistor

υs

Dependent voltage source

Path: Continuous sequence of branches with no node encountered more than once. Extraordinary path: Path between two adjacent extraordinary nodes.

6A

+
0 p 0, the device is a recipient of power. As we know, the law of conservation of power requires that if the device receives 2.4 W of power, then the battery has to deliver exactly that same amount of power. For the battery, the current entering its (+) terminal is −0.2 A (because 0.2 A of current is shown leaving that terminal), so according to the passive sign convention, the power that would be absorbed by the battery (had it been a passive device) is

18 V 12 V

+ _

6V

(b) For device 1 in Fig. 1-23(b), the current entering its (+) terminal is 3 A. Hence, P1 = V1 I1 = 18 × 3 = 54 W, and the device is a power recipient. For device 2,

Device 2

(b)

Pbat = 12(−0.2) = −2.4 W. The fact that Pbat is negative is confirmation that the battery is indeed a supplier of power.

3A

Figure 1-23: Circuits for Example 1-5.

Example 1-6: Energy Consumption

A resistor connected to a 100 V dc power supply was consuming 20 W of power until the switch was turned off, after which the voltage decayed exponentially to zero. If t = 0 is defined as the time at which the switch was turned to the off position and if the subsequent voltage variation is given by υ(t) = 100e−2t V

for t ≥ 0

P2 = V2 I2 = (−6) × 3 = −18 W,

(where t is in seconds), determine the total amount of energy consumed by the resistor after the switch was turned off.

and the device is a supplier of power (because P2 is negative). By way of confirmation, the power associated with the battery is

Solution: Before t = 0, the current flowing through the resistor was I = P /V = 20/100 = 0.2 A. Hence, the resistance R of the resistor is R=

Pbat = 12(−3) = −36 W,

100 V = = 500 , I 0.2

and the current variation after the switch was turned off is thereby satisfying the law of conservation of power, which requires the net power of the overall circuit to be exactly zero.

i(t) =

υ(t) = 0.2e−2t A R

for t ≥ 0.

“book” — 2015/5/4 — 6:55 — page 35 — #35

1-6

CIRCUIT ELEMENTS

35

The instantaneous power is p(t) = υ(t) · i(t) = (100e−2t )(0.2e−2t ) = 20e−4t W. We note that the power decays at a rate (e−4t ) much faster than the rate for current and voltage (e−2t ). The total energy dissipated in the resistor after engaging the switch is obtained by integrating p(t) from t = 0 to infinity (the integral equation form of Eq. (1.8)), namely ∞

∞ p(t) dt =

W = 0

20e−4t dt = −

20 −4t ∞ e  = 5 J. 0 4

0

1-6.1

Concept Question 1-10: Explain how node voltage relates to voltage difference. To what do the (+) and (−) leads of the voltmeter connect to in each case? (See )

Exercise 1-6: If a positive current is flowing through a resistor from its terminal a to its terminal b, is υab positive or negative? Answer: υab > 0. (See

C3

)

Exercise 1-7: A certain device has a voltage difference of 5 V across it. If 2 A of current is flowing through it from its (−) voltage terminal to its (+) terminal, is the device a power supplier or a power recipient, and how much energy does it supply or receive in 1 hour? Answer: P = V I = 5(−2) = −10 W. Hence, the

device is a power supplier. Since p(t) = (not timeC) varying), |W | = |P | t = 36 kJ. (See

Exercise 1-8: A car radio draws 0.5 A of dc current when connected to a 12 V battery. How long does it take for the radio to consume 1.44 kJ? Answer: 4 minutes. (See

with idealized characteristics. The equivalent circuit offers a circuit behavior that closely resembles the behavior of the actual electronic circuit or device over a certain range of specified conditions, such as the range of input signal level or output load resistance. The set of basic elements commonly used in circuit analysis include voltage and current sources; passive elements (which include resistors, capacitors, and inductors); and various types of switches. The basic attributes of switches were covered in Section 1-5.1. The nomenclature and current– voltage relationships associated with the other two groups are the subject of this section.

)

1-6 Circuit Elements Electronic circuits used in functional systems employ a wide range of circuit elements, including transistors and integrated circuits. The operation of most electronic circuits and devices— no matter how complex—can be modeled (represented) in terms of an equivalent circuit composed of basic elements

i–υ Relationship

The relationship between the current flowing through a device and the voltage across it defines the fundamental operation of that device. As was stated earlier, Ohm’s law states that the current i entering into the (+) terminal of the voltage υ across a resistor is given by υ i= . R This is called the i–υ relationship for the resistor. We note that the resistor exhibits a linear i–υ relationship, meaning that i and υ always vary in a proportional manner, as shown in Fig. 1-24(a), so long as R remains constant. A circuit composed exclusively of elements with linear i–υ responses is called a linear circuit. The linearity property of a circuit is an underlying requirement for the various circuit analysis techniques presented in this and future chapters. Diodes and transistors exhibit nonlinear i–υ relationships. To apply the analysis techniques specific to linear circuits to circuits containing nonlinear devices, we can represent those devices in terms of linear subcircuits that contain dependent sources. The concept of a dependent source and how it is used is introduced in Section 1-6.4.

1-6.2

Independent Voltage Source

An ideal, independent voltage source provides a specified voltage across its terminals, independent of the type of load or circuit connected to it (so long as it is not a short circuit). Hence, for a voltage source with a specified voltage Vs , its i–υ relationship is given by υ = Vs

for any i = ∞.

The i–υ profile of an ideal voltage source is a vertical line, as illustrated in Fig. 1-24(b).

“book” — 2015/5/4 — 6:55 — page 36 — #36

36

CHAPTER 1

i

υ i= R

CIRCUIT TERMINOLOGY

Rs Resistor 1 Slope = R

υs

+ _

RL

υ

Realistic voltage source Load (a) Realistic voltage source connected to load RL

(a) i

υ = Vs Ideal voltage source

Is Vs

i = Is Ideal current source

Rs

RL

υ

Realistic current source Load (b) Realistic current source connected to load RL

(b)

Figure 1-25: (a) A realistic voltage source has a nonzero series resistance Rs , which can be replaced with a short circuit if Rs is much smaller than the load resistance RL . (b) A realistic current source has a nonzero parallel resistance Rs , which can be replaced with an open circuit if Rs  RL .

υs υs = αυx

is

VCVS Slope = α υx

(c)

Figure 1-24: i–υ relationships for (a) an ideal resistor, (b) ideal, independent current and voltage sources, and (c) a dependent, voltage-controlled voltage source (VCVS).

The circuit symbol used for independent sources is a circle, as shown in Table 1-5, although for dc voltage sources the traditional “battery” symbol is used as well. A household electrical outlet connected through an electrical power-distribution network to a hydroelectric- or nuclearpower generating station provides continuous power at an approximately constant voltage level. Hence, it may be classified appropriately as an independent voltage source. On a shorter time scale, a flashlight’s 9-volt battery may be regarded as an independent voltage source, but only until its stored charge has been used up by the light bulb. Thus, strictly speaking, a battery is a storage device (not a generator), but we tend to treat it as a generator so long as it acts like a constant voltage source. In reality, no sources can provide the performance specifications ascribed to ideal sources. If a 5 V voltage source

is connected across a short circuit, for example, we run into a serious problem of ambiguity. From the standpoint of the source, the voltage is 5 V, but by definition, the voltage across the short circuit is zero. How can it be both zero and 5 V simultaneously? The answer resides in the fact that our description of the ideal voltage source breaks down in this situation. Most often, in such cases, the circuit malfunctions as well. Short-circuiting a battery will draw more current than the battery is intended to provide, thereby overheating it, damaging it, and possibly causing a fire or explosion.  A more realistic model for a voltage source includes an internal series resistor, as shown in Fig. 1-25(a).  A real voltage source (which may have an elaborate circuit configuration) behaves like a combination of an equivalent, ideal voltage source υs in series with an equivalent resistance Rs . Usually, the voltage source is designed such that its series resistance Rs is much smaller than the resistance values of the types of loads it is intended to energize. Under such a condition, Rs becomes inconsequential and can be ignored, in which case the realistic voltage source behaves approximately the same as an ideal voltage source.

“book” — 2015/5/4 — 6:55 — page 37 — #37

1-6

CIRCUIT ELEMENTS

37

Table 1-5: Voltage and current sources.

Independent Sources Ideal Voltage Source

Realistic Voltage Source

Rs Vs

+ -

Vs

or

Battery

+

+_ −

υs

dc source

+

+_ −

Realistic Current Source

is

is

dc source

+

+ -_ Any source

Any source*

Ideal Current Source

Is

υs

Any source

Rs

Any source

Dependent Sources Voltage-Controlled Voltage Source (VCVS)

+
0.

R1

+ −

V1 4

Figure E1.9 Answer: Ix = 2.5 A. (See

V0

Ix =



I

SPDT t=0

R3 R6

R5

R7 (c) t > 5 s

Figure 1-30: Solutions for circuit in Fig. 1-29.

12 V

+ _ 3Ω

R4



Figure E1.10 Answer: (a) I = 4 A, (b) I = 3 A. (See

)

“book” — 2015/5/4 — 6:55 — page 41 — #41

SUMMARY

41

Summary Concepts • Active devices (such as transistors and ICs) require an external power source to operate; in contrast, passive devices (resistors, capacitors, and inductors) do not. • Analysis and synthesis (design) are complementary processes. • Current is related to charge by i = dq/dt; voltage between locations a and b is υab = dw/dq, where dw is the work (energy) required to move dq from b to a; and power p = υi. • Passive sign convention assigns i direction as entering

Mathematical and Physical Models Ohm’s law

the (+) side of υ; if p > 0, the device is recipient (consumer) of power, and if p < 0, it is a supplier of power. • Node voltage refers to the voltage difference between the node and ground by selecting Vground = 0. • Independent voltage and current sources are real sources of energy; dependent sources are artificial representations used in modeling the nonlinear behavior of active devices (transistors and integrated circuits) in terms of an equivalent linear circuit.

Passive sign convention

Current i = dq/dt Direction of i = direction of flow of (+) charge

p>0 p 2 s

Figure P1.38: Circuit for Problem 1.38.

R1 V = 2Ix _

10 Ω

+

V0

20 Ω

10 V

Ix

+ _

R2

R3 R4

+ _

t=2s R5

+ _



t=0

R6

15 V Figure P1.42: Circuit for Problem 1.42.

30 Ω Figure P1.39: Circuit for Problem 1.39.

Vx

+ 2.5 A

2Ω Iy = 0.1Vx

(a) t < 0

4Ω 2Ω

(b) 0 < t < 2 s

+ _

+ Vz 5 Ω _

1.43 For the circuit in Fig. P1.43, generate circuit diagrams that include only those elements that have current flowing through them for

19 V

(c) t > 2 s

Figure P1.40: Circuit for Problem 1.40.



R1

6V

V1

+ _



+ V1 _ 6A



Ix =

V1 2

+ V2 _

SPST t=0

R3

SPDT t=2s

R2 1 2 R 5 t=0

R6 SPST

Figure P1.41: Circuit for Problem 1.41.

Figure P1.43: Circuit for Problem 1.43.

R4

“book” — 2015/5/4 — 6:55 — page 49 — #49

PROBLEMS

49

1.44 The switch in the circuit of Fig. P1.44 closes at t = 0. Which elements are in series and which are in parallel at (a) t < 0 and (b) t > 0?

R1 + _

υs 3

1 R2 2

R3 t=0

R5

Figure P1.44: Circuit for Problem 1.44.

R4 4 R6

Potpourri Questions 1.45 What aspect of electrical engineering particularly interests you? Check out http://spectrum.ieee.org/ to learn more. 1.46 Will the prediction of Moore’s Law continue to hold true indefinitely? If not, why not? 1.47 Provide a definition of what the term “nanotechnology” means to you. 1.48 What is the typical voltage level associated with lightning? With a bird standing on a power line (foot to foot)?

“book” — 2015/5/4 — 6:57 — page 50 — #1

2

CHAPTER

Resistive Circuits Contents 2-1 TB3 2-2 2-3 TB4 2-4 2-5 2-6 TB5 2-7

Overview, 51 Ohm’s Law, 51 Superconductivity, 57 Kirchhoff’s Laws, 60 Equivalent Circuits, 67 Resistive Sensors, 70 Wye–Delta (Y–) Transformation, 80 The Wheatstone Bridge, 84 Application Note: Linear versus Nonlinear i–υ Relationships, 86 Light-Emitting Diodes (LEDs), 90 Introducing Multisim, 94 Summary, 100 Problems, 101

Objectives Learn to: 

Apply Ohm’s law and explain the basic properties of piezoresistivity and superconductivity.



State Kirchhoff’s current and voltage laws; apply them to resistive circuits.



Define circuit equivalency, combine resistors in series and in parallel, and apply voltage and current division.



Apply source transformation between voltage and current sources and Y– circuits.



Describe the operation of the Wheatstone-bridge circuit and how it is used to measure small deviations.



Use Multisim and myDAQ to analyze simple circuits.

Microfabricated pressure sensor The basic laws of circuit theory are used to develop fluency in analyzing resistive circuits and characterizing their performance.

“book” — 2015/5/4 — 6:57 — page 51 — #2

2-1

OHM’S LAW

51

Overview

Table 2-1: Conductivity and resistivity of some common materials at 20 ◦ C.

The study of any field of inquiry starts with nomenclature: defining the terms specific to that field. That is exactly what we did in the preceding chapter. We introduced and defined electric current, voltage, power, open and closed circuits, and dependent and independent voltage and current sources, among others. Now, we are ready to acquire our first set of circuitanalysis tools, which will enable us to analyze a variety of different types of circuits. We limit our discussion to resistive circuits, namely those circuits containing only sources and resistors. (In future chapters, we will extend these tools to circuits containing capacitors, inductors, and other elements.) Our new toolbox includes three simple, yet powerful laws— Ohm’s law and Kirchhoff’s voltage and current laws—and several circuit simplification and transformation techniques. You will learn how to divide the voltage (using voltage dividers) and current (using current dividers), how to combine resistors in series and parallel combinations, how to analyze resistive sensors using Wheatstone bridges, how to use diodes to control the direction of a current, plus how to use a light-emitting diode (LED) as a visual output, warning light, etc. You will also learn how to use Multisim to simulate and analyze your circuits, and how to build a circuit on a circuit board and measure its properties using your computer via the NI myDAQ.

2-1 Ohm’s Law  The conductivity σ of a material is a measure of how easily electrons can drift through the material when an external voltage is applied across it. Resistivity (ρ) is the inverse (1/σ ) of conductivity.  Materials are classified as conductors (primarily metals), semiconductors, or dielectrics (insulators) according to the magnitudes of their conductivities. Tabulated values of σ expressed in units of siemens per meter (S/m) are given in Table 2-1 for a select group of materials. The siemen is the inverse of the ohm, S = 1/ , and the inverse of σ is called the resistivity ρ, ρ=

1 σ

(-m),

(2.1)

which is a measure of how well a material impedes the flow of current through it. The conductivity of most metals is on the order of 107 S/m, which is 17 or more orders of magnitude

Conductivity σ (S/m)

Resistivity ρ (-m)

Conductors Silver Copper Gold Aluminum Iron Mercury (liquid)

6.17 × 107 5.81 × 107 4.10 × 107 3.82 × 107 1.03 × 107 1.04 × 106

1.62 × 10−8 1.72 × 10−8 2.44 × 10−8 2.62 × 10−8 9.71 × 10−8 9.58 × 10−7

Semiconductors Carbon (graphite) Pure germanium Pure silicon

7.14 × 104 2.13 4.35 × 10−4

1.40 × 10−5 0.47 2.30 × 103

Insulators Paper Glass Teflon Porcelain Mica Polystyrene Fused quartz

∼ 10−10 ∼ 10−12 ∼ 3.3 × 10−13 ∼ 10−14 ∼ 10−15 ∼ 10−16 ∼ 10−17

∼ 1010 ∼ 1012 ∼ 3 × 1012 ∼ 1014 ∼ 1015 ∼ 1016 ∼ 1017

Common materials Distilled water Drinking water Sea water Graphite Rubber

5.5 × 10−6 ∼ 5 × 10−3 4.8 1.4 × 10−5 1 × 10−13

1.8 × 105 ∼ 200 0.2 71.4 × 103 1 × 1013

Biological tissues Blood Muscle Fat

∼ 1.5 ∼ 1.5 ∼ 0.1

∼ 0.67 ∼ 0.67 10

Material

greater than the conductivity of typical insulators. Common semiconductors, such as silicon and germanium, fall in the inbetween range on the conductivity scale. The values of σ and ρ given in Table 2-1 are specific to room temperature at 20 ◦ C. In general, the conductivity of a metal increases with decreasing temperature. At very low temperatures (in the neighborhood of absolute zero), some conductors become superconductors, because their conductivities become practically infinite and their corresponding resistivities approach zero. To learn more about superconductivity, refer to Technology Brief 3.

“book” — 2015/5/4 — 6:57 — page 52 — #3

52

CHAPTER 2

Table 2-2: Diameter d of wires, according to the American

l

Wire Gauge (AWG) system.

σ

R=

A

l σA

Figure 2-1: Longitudinal resistor of conductivity σ , length , and cross-sectional area A.

2-1.1

Resistance

 The resistance R of a device incorporates two factors: (a) the inherent bulk property of its material to conduct (or impede) current, represented by the conductivity σ (or resistivity ρ), and (b) the shape and size of the device.  For a longitudinal resistor (Fig. 2-1), R is given by

R=

  =ρ σA A

(),

(2.2)

where  is the length of the device and A is its cross-sectional area. In addition to its direct dependence on the resistivity ρ, R is directly proportional to , which is the length of the path that the current has to flow through, and inversely proportional to A, because the larger A is, the more electrons can drift through the material. Every element of an electric circuit has a certain resistance associated with it. This even includes the wires used to connect devices to each other, but we usually treat them like zeroresistance segments because their resistances are so much smaller than the resistances of the other devices in the circuit. To illustrate with an example, let us consider a 10 cm long segment of one of the wire sizes commonly found in circuit boards, such as the AWG-18 copper wire. According to Table 2-2, which lists the diameter d for various wire sizes as specified by the American Wire Gauge (AWG) system, the AWG-18 wire has a diameter d = 1 mm. Using the values specified for  and d and the value for ρ of copper given in Table 2-1, we have R=ρ

RESISTIVE CIRCUITS

 0.1  = 1.72 × 10−8 × =ρ A π(d/2)2 π(0.5 × 10−3 )2 = 2.2 × 10−3  = 2.2 m.

AWG Size Designation

Diameter d (mm)

0 2 4 6 10 14 18 20

8.3 6.5 5.2 4.1 2.6 1.6 1.0 0.8

 Thus, R of a 10 cm long AWG-18 copper wire is on the order of milliohms. If the wire segment connects to circuit elements with resistances of ohms or larger, ignoring the resistance of the wire would have no significant impact on the overall behavior of the circuit.  The preceding justification should be treated with some degree of caution. While it is true that a piece of wire may be treated like a short circuit in the majority of circuit configurations, there are certain situations for which such an assumption may not be valid. One obvious example is when the wire is very long, as in the case of a kilometers long electric power-transmission cable. Another is when very thin wires or channels with micron-size diameters are used in microfabricated circuits. Resistive elements used in electronic circuits are fabricated in many different sizes and shapes to suit the intended application and requisite circuit architecture. Discrete resistors usually are cylindrical in shape and made of a carbon composite. Hybrid and miniaturized circuits use film-shaped metal or carbon resistors. In integrated circuits, resistive elements are fabricated through a diffusion process (see Technology Brief 7). Figure 2-2 displays photographs of three types of resistors, amongst which the tubular-shaped resistor is the most familiar. Resistors are generally marked with a banded color code to denote the resistor’s specifications: (a) 4-Band color code: b1

b2

b4

b5

Note that a wider spacing exists between b4 and b5 than between the earlier bands. The resistor value is given by R = (b1 b2 ) × 10b4 ± b5 , with the values of b1 , b2 , b4 , and b5 specified by the color code shown in Fig. 2-2. For example, = 25 × 100 ± 10% = 25 ± 10% .

“book” — 2015/5/4 — 6:57 — page 53 — #4

2-1

OHM’S LAW

53

Rotatable-shaft potentiometer Rotating dial Resistive material

Screw-top potentiometer

1

4 1 2 k ± 1% 5 ppm /˚C 1

Rmax

2

3

Movable wiper R 13 3 R23

Potentiometer resistor 2 4 bands

25 Ω, 10%

5 bands

62 MΩ, 5%

6 bands

500 kΩ, 0.25%, 15 ppm

Silver b1

Gold

b2

b3

Potentiometer

b4 0.01

b5 10%

0.1

5%

Black

0

0

0

1

Brown

1

1

1

10

1%

b6 100 ppm

Red

2

2

2

100

2%

50 ppm

Orange

3

3

3

1K

15 ppm

Yellow

4

4

4

10K

25 ppm

Green

5

5

5

100K

0.5%

Blue

6

6

6

1M

0.25%

10 ppm

Purple

7

7

7

10M

0.1%

5 ppm

Gray

8

8

8

White

9 1st digit

Multiplier Tolerance Temperature 9 9 × 10b4 coefficient 2nd digit 3rd digit # of zeros ppm /˚C 4-, 5-, and 6-band color code system

Figure 2-2: Various types of resistors. Tubular-shaped resistors usually are color-coded by 4-, 5-, or 6-band systems. (b) 5-Band color code: b1

b2

b3

b4

In this case R = (b1 b2 b3 ) × 10b4 ± b5 .

b5

(c) 6-Band color code: b1

b2

b3

b4

b5

b6

This code adds one more piece of information in the form of b6 which denotes the temperature coefficient of the resistor, measured in parts-per-million/ ◦ C.

“book” — 2015/5/4 — 6:57 — page 54 — #5

54

CHAPTER 2

Table 2-3: Common resistor terminology. Thermistor Piezoresistor Light-dependent R (LDR) Rheostat Potentiometer

R sensitive to temperature R sensitive to pressure R sensitive to light intensity 2-terminal variable resistor 3-terminal variable resistor

For some metal oxides, the resistivity ρ exhibits a strong sensitivity to temperature. A resistor manufactured of such materials is called a thermistor (Table 2-3), and it is used for temperature measurement, temperature compensation, and related applications. Another interesting type of resistor is the piezoresistor, which is used as a pressure sensor in many household appliances, automotive systems, and biomedical devices. More coverage on resistive sensors is available in Technology Brief 4. Certain applications, such as volume adjustment on a radio, may call for the use of a resistor with variable resistance. The rheostat and the potentiometer are two standard types of variable resistors in common use. The rheostat [Fig. 2-3(a)] is a twoterminal device with one of its terminals connected to one end of a resistive track and the other terminal connected to a movable wiper. Movement of the wiper across the resistive track, through rotation of a shaft, can change the resistance between the two terminals from (theoretically) zero resistance to the full resistance value of the track. Thus, if the total resistance of the track is Rmax , the rheostat can provide any resistance between zero and Rmax .

The potentiometer is a three-terminal device. Terminals 1 and 2 in Fig. 2-3(b) are connected to the two ends of the track (with total resistance Rmax ) and terminal 3 is connected to a movable wiper. When terminal 3 is at the end next to terminal 1, the resistance between terminals 1 and 3 is zero and that between terminals 2 and 3 is Rmax . Moving terminal 3 away from terminal 1 increases the resistance between terminals 1 and 3 and decreases the resistance between terminals 2 and 3. A potentiometer can be used as a rheostat by connecting to only terminals 1 and 3.

2-1.2

i–υ Characteristics of Ideal Resistor

Based on the results of his experiments on the nature of conduction in circuits, German physicist Georg Simon Ohm (1787–1854) formulated in 1826 the i–υ relationship for a resistor, which has become known as Ohm’s law. He discovered that the voltage υ across a resistor is directly proportional to the current i flowing through it, namely υ = iR,

Movable wiper

1

R Rmax

Rmax

Movable wiper R 13 3 R23

2 Terminal 2 (a) Rheostat

(b) Potentiometer

Figure 2-3: (a) A rheostat is used to set the resistance between terminals 1 and 2 at any value between zero and Rmax ; (b) the wiper in a potentiometer divides the resistance Rmax among R13 and R23 .

(2.3)

with the resistance R being the proportionality factor.  In compliance with the passive sign convention, current enters a resistor at the “+” side of the voltage across it. υab = υa − υb

υa

R

+

i i=

Terminal 1

RESISTIVE CIRCUITS

υa − υb R

_

υb 

An ideal linear resistor is one whose resistance R is constant and independent of the magnitude of the current flowing through it, in which case its i–υ response is a straight line (Fig. 2-4(a)). In practice, the i–υ response of a real linear resistor is indeed approximately linear, as illustrated in Fig. 2-4(b), so long as i remains within the linear region defined by −imax to imax . The slope of the curve is the resistance R. Outside this range, the response deviates from the straight-line model. When we use Ohm’s law as expressed by Eq. (2.3), we tacitly assume that the resistor is being used in its linear range of operation. Some resistive devices exhibit highly nonlinear i–υ characteristics. These include diode elements and light-bulb filaments, among others. Unless noted otherwise, the common use of the term resistor in circuit analysis and design usually refers to the linear resistor exclusively.

“book” — 2015/5/4 — 6:57 — page 55 — #6

2-1

OHM’S LAW

55

υ 2V

R1

+

R = 2 kΩ

R=∞

V1 _ i

R = 1 kΩ

1V

R=0

+ υ_

R2

R3

I

A

(a) Same current flows through all elements

R

R4

i

+

V2 _

0.5 mA 1 mA

(a) Ideal resistor

R5 V

υ

com (b) Same voltage exists across R4 and R5 Figure 2-5: In-series and in-parallel connections.

−imax imax

i

Current-limiting devices, such as fuses and circuit breakers, are used to protect against dangerous overloading of circuits.

Linear region (b) Real resistor

2-1.3

Figure 2-4: i–υ responses of ideal and real resistors.

The flow of current in a resistor leads to power dissipation in the form of heat (or the combination of heat and light in the case of a light bulb’s filament). Using Eq. (2.3) in Eq. (1.9) provides the following expression for the power p dissipated in a resistor:

p = iυ = i 2 R =

υ2 R

(W).

(2.4)

In-Series and In-Parallel Connections

Recall from Chapter 1 that two or more elements are considered to be connected in series if the same current flows through all of them. This is indeed the case for voltage source V1 and the resistors shown in Fig. 2-5(a). For two or more elements to be in parallel, they have to share the same voltage, which is the case for R4 and R5 in Fig. 2-5(b).

Example 2-1: Series Connection Resistances for a dc Motor

 The power rating of a resistor defines the maximum continuous power level that the resistor can dissipate without getting damaged. Excessive heat can cause melting, smoke, and even fire. 

A 12 V car battery is connected via a 6 m long, twin-wire cable to a dc motor that drives the wiper blade on the rear window. The cable is copper AWG-10 and the motor exhibits to the rest of the circuit an equivalent resistance Rm = 2 . Determine: (a) the resistance of the cable and (b) the fraction of the power contributed by the battery that gets delivered to the motor.

For electric circuits with a fixed voltage (such as a 120 V for a house), the power rating refers to the maximum current limit.

Solution: The circuit described in the problem statement is represented by Fig. 2-6.

“book” — 2015/5/4 — 6:57 — page 56 — #7

56

CHAPTER 2

Wire (6 m long) Rtop

+ 12 V _

Rm (motor resistance)

Wire Rbottom

Car battery

i

υs(t)

Same as positive terminal of υs(t) i1

+ _

RESISTIVE CIRCUITS

R1 1 kΩ

i2 R2

500 Ω

i3 R3 250 Ω

Rc = resistance of both wires Same as negative terminal of υs(t)

Figure 2-6: Circuit for Example 2-1.

Figure 2-7: Circuit for Example 2-2. (a) We need to include both the top wire and the bottom wire, as each represents a resistor through which the current flows, and therefore contributes to the resistive losses of the circuit. With  = 12 m (total for twin wires), ρ = 1.72 × 10−8 -m for copper, A = π(d/2)2 , and d = 2.6 mm for AWG-10, the cable resistance Rc is Rc = ρ

 12 = 0.04 . = 1.72 × 10−8 × A π(1.3 × 10−3 )2

(b) The total resistance in the circuit is equal to the sum of the cable and motor resistances. [In a later section, we will learn that the resistance of two resistors connected in series is simply equal to the sum of their resistances.] Hence, R = Rc + Rm = 0.04 + 2 = 2.04 . Consequently, the current flowing through the circuit is 12 V = = 5.88 A, I= R 2.04 and the power contributed by the battery P and the power delivered to the motor Pm are: P = I V = 5.88 × 12 = 70.56 W and Pm = I 2 Rm = (5.88)2 × 2 = 69.15 W, and the fraction of P delivered to the load (motor) is Fraction =

69.15 Pm = = 0.98 or 98 percent. P 70.56

Thus, 2 percent of the power is dissipated in the cable. Concept Question 2-1: If the terminals of the battery in Fig. 2-6 were corroded, how would that change the problem and the results? (See )

Example 2-2: Parallel Loads

Three loads—a 1 k light bulb, a 500  computer, and a 250  TV, each represented by a resistor, are connected in parallel to a household ac voltage source as shown in Fig. 2-7. The source is cosinusoidal in time at a frequency of 60 Hz and its amplitude is 170 V. Hence, it can be described as υs (t) = 170 cos(2π × 60t) = 170 cos(377t) V. Determine the currents supplied by the source to the three loads. Solution: All three loads share the same positive terminal (node) of υs (t) on one end and the same negative terminal (node) on the other. Consequently, application of Ohm’s law leads to υs (t) 170 = 3 cos(377t) = 0.17 cos(377t) A, R1 10 υs (t) 170 i2 (t) = cos(377t) = 0.34 cos(377t) A, = R2 500 υs (t) 170 i3 (t) = cos(377t) = 0.68 cos(377t) A. = R3 250 i1 (t) =

As we see in the next section, the current i supplied by the source is the sum of the three load currents, i(t) = i1 + i2 + i3 = 1.19 cos(377t) A.

Concept Question 2-2: How does the magnitude of the conductivity of a metal, such as copper, compare with that of a typical insulator, such as mica? (See )

Concept Question 2-3: What is piezoresistivity, and how

is it used? (See

)

“book” — 2015/5/4 — 6:57 — page 57 — #8

TECHNOLOGY BRIEF 3: SUPERCONDUCTIVITY

57

Technology Brief 3 Superconductivity When an electric voltage is applied across two points in a conductor, such as copper or silver, current flows between them. The relationship between the voltage difference V and the current I is given by Ohm’s law, V = IR, where R is the resistance of the conducting material between the two points. It is helpful to visualize the electric current as a fluid of electrons flowing through a dense forest of sturdy metal atoms, called the lattice. Under the influence of the electric field (induced by the applied voltage), the electrons can attain very high instantaneous velocities, but their overall forward progress is impeded by the frequent collisions with the lattice atoms. Every time an electron collides and bounces off an atom, some of that electron’s kinetic energy is transferred to the atom, causing the atom to vibrate—which heats up the material—and causing the electron to slow down. The resistance R is a measure of how much of an obstacle the resistor poses to the flow of current, as well as a measure of how much heat it generates for a given current. The power dissipated in R is I 2R if I is a dc current, and it is I 2R if the current is ac with an amplitude I . Can a conductor ever have zero resistance? The answer is most definitely yes! In 1911, the Dutch physicist Heike Kamerlingh Onnes developed a refrigeration technique so powerful that it could cool helium down low enough to condense it into liquid form at 4.2 K (0 kelvin = −273 ◦C). Into his new liquid helium container, he immersed (among other things) mercury; he soon discovered that the resistance of a solid piece of mercury at 4.2 K was zero! The phenomenon, which was completely unexpected and not predicted by classical physics, was coined superconductivity. According to quantum physics, many materials experience an abrupt change in behavior (called a phase transition) when cooled below a certain critical temperature TC. Superconductors have some amazing properties. The current in a superconductor can persist with no external voltage applied. Even more interesting, currents have been observed to persist in superconductors for many years without decaying. When a magnet is brought close to the surface of a superconductor, the currents induced by the magnetic field are mirrored exactly by the superconductor (because the superconductor’s resistance is zero), and consequently the magnet is repelled (Fig. TF3-1). This property has been used to demonstrate magnetic levitation and is the basis of some super-fast maglev trains (Fig. TF3-2) being

Figure TF3-1: The Meissner effect, or strong diamagnetism, seen between a high-temperature superconductor and a rare earth magnet. (Courtesy of Pacific Northwest National Laboratory.)

1 2

Figure TF3-2: Maglev train. (Courtesy of Central Japan Railway Company.)

used around the world. The same phenomenon is used in the Magnetic Resonance Imaging (MRI) machines that hospitals use to perform 3-D scans of organs and tissues (Fig. TF3-3) and in Superconducting Quantum Interference Devices (SQUIDs) to examine brain activity at high resolution.

“book” — 2015/5/4 — 6:57 — page 58 — #9

58

TECHNOLOGY BRIEF 3: SUPERCONDUCTIVITY

Figure TF3-3: Magnetic Resonance Imaging machine. (Courtesy GE Healthcare.)

Superconductivity is one of the last frontiers in solidstate physics (see Table TT3-1). Even though the physics of low-temperature superconductors (like mercury, lead, niobium nitride, and others) is now fairly well understood, a different class of high-temperature superconductors still defies complete theoretical explanation. This class of materials was discovered in 1986 when Alex Muller ¨ and Georg Bednorz, at IBM Research Laboratory in Switzerland, created a ceramic compound that superconducted at 30 K. This discovery was followed by the discovery of other ceramics with even higher TC values; the now-famous YBCO ceramic discovered at the University of Alabama-Huntsville (1987) has

a TC of 92 K, and the world record holder is a group of mercury-cuprate compounds with a TC of 138 K (1993). New superconducting materials and conditions are still being found; carbon nanotubes, for example, were recently shown to have a TC of 15 K (Hong Kong University, 2001). Are there highertemperature superconductors? What theory will explain this higher-temperature phenomenon? Can so-called room-temperature superconductors exist? For engineers (like you) the challenges are just beginning: How can these materials be made into useful circuits, devices, and machines? What new designs will emerge? The race is on!

Table TT3-1: Critical temperatures. Critical Temperature Tc [K]

Material

Type

138 138 92

HgBa2 Ca2 Cu3 Ox Bi2 Sr2 Ca2 Cu3 O10 (BSCCO) YBa2 Cu3 O7 (YBCO)

Copper-oxide superconductors

55 41 26

SmFeAs CeFeAs LaFeAs

Iron-based superconductors

18 10 9.2 4.2

Nb3 Sn NbTi Nb Hg (mercury)

Metallic low-temperature superconductors

“book” — 2015/5/4 — 6:57 — page 59 — #10

2-1

OHM’S LAW

59

i–υ Characteristics of LEDs

Concept Question 2-4: What is meant by the linear

2-1.4

region of a resistor? Is it related to its power rating? (See )

A resistor is a bidirectional device, meaning that current can flow through it in either direction. This is because it is constructed of the same material along the dimension between its two terminals. In contrast, a diode allows current to flow in only one direction. It is built of two sections of different semiconductor materials, denoted p and n in Fig. 2-8(a). The p-type material has excess positive charges and the n-type material has excess negative charges. When connected to a voltage source, the diode acts like a resistor in one direction, but like an open circuit in the other. Specifically: (a) Reverse bias: If the voltage VD applied across the diode is negative (relative to its own terminals), as shown in Fig. 2-8(b), no current flows through it, which is equivalent to having infinite resistance. That is, the diode behaves like an open circuit. (b) Forward bias: If the voltage VD is positive, as in Fig. 2-8(c), current will flow through the diode, but the relationship between I and VD is not a constant. For a

Exercise 2-1: A cylindrical resistor made of carbon has a power rating of 2 W. If its length is 10 cm and its circular cross section has a diameter of 1 mm, what is the maximum current that can flow through the resistor without damaging it? Answer: 1.06 A. (See

)

Exercise 2-2: A rectangular bar made of aluminum has a current of 3 A flowing through it along its length. If its length is 2.5 m and its square cross section has 1 cm sides, how much power is dissipated in the bar at 20 ◦C? Answer: 5.9 mW. (See

)

Anode

I=0

+ VD _

_ +

I > 0 if VD ≥ VF

+ VD _

No conduction, diode like open circuit

+ _

Diode like nonlinear resistor

p

Cathode

Red Amber

Forward current I (mA)

Green Blue

n

(c) Forward-biased diode Yellow

(b) Reverse-biased diode

50

(a) Diode 40

(d) i-υ plots for LEDs

30 LED ON 20 LED OFF

+ _VF RD (e) LED equivalent circuit

10 0

0

1

1.6 2 VF (red)

3 3.3 4 VF (green)

5

VD

Figure 2-8: p-n junction diode (a) configuration, (b) reverse biased, (c) forward biased, (d) typical i-υ plots for LEDs, and (e) LED equivalent circuit.

“book” — 2015/5/4 — 6:57 — page 60 — #11

60

CHAPTER 2

resistor, VD /I = R and R is a constant, but for a diode the relationship between VD and I is more complicated. However, its i–υ relationship can be approximated by I = aVD2

G=

Exercise 2-3: A certain type of diode exhibits a nonlinear relationship between υ—the voltage across it—and i— the current entering into its (+) voltage terminal. Over its operational voltage range (0 to 1 V), the current is given by

for 0 ≤ υ ≤ 1 V.

Answer: R =

2 , υ

R ∞ 200  20  4 2

(See

)

1 R

(S),

(2.5)

and its unit is −1 , which is called the siemen (S, or sometimes called “mho”). In terms of G, Ohm’s law can be rewritten in the form υ i= = Gυ, (2.6) R and the expression for power becomes p = iυ = Gυ 2

(W).

(2.7)

Since G = 1/R, what is the point in dealing with both G and R? The answer is: convenience. In some circuit solutions it is easier to work with R for all resistors in the circuit, whereas in other circuit configurations (especially those in parallel) it may be easier to work with conductances instead.

2-2

Kirchhoff’s Laws

Circuit theory—encompassing both analysis and synthesis— is built upon a foundation comprised of a small number of fundamental laws. Among the cornerstones are Kirchhoff’s current and voltage laws. Kirchhoff’s laws, which constitute the subject of this section, were introduced by the German physicist Gustav Robert Kirchhoff (1824–1887) in 1847, some 21 years after a fellow German, Georg Simon Ohm, developed his famous law.

2-2.1

Determine how the diode’s effective resistance varies with υ and calculate its value at υ = 0, 0.01 V, 0.1 V, 0.5 V, and 1 V. υ 0 0.01 V 0.1 V 0.5 V 1V

Conductance

The reciprocal of resistance is called conductance,

(VD > 0),

where a is a constant that depends on the semiconductor material used to build the LED. A light-emitting diode is a special kind of diode in that it emits light if I exceeds a certain threshold. Figure 2-8(d) displays plots of I versus VD for five LEDs of different colors. The color of light emitted by an LED depends on the semiconductor compounds from which it is constructed. The voltage at which the diode becomes approximately linear is the forward bias voltage VF , and it becomes part of the diode model shown in Fig. 2-8(e). For the typical family of LEDs shown in Fig. 2-8(d), the current I has to exceed 20 mA in order for the LED to fully light up. This current threshold has a corresponding voltage threshold called the forward voltage VF . Below this threshold, the diode conducts little or no current and is considered “OFF” (although it does generate a small amount of light). For the red LED, for example, VF = 1.6 V, and the current flowing through the LED at that voltage is exactly 20 mA. Higher values of VF are required to cause the LEDs of the other colors to emit light. When we analyze a circuit containing an LED, the LED can be modeled as an ideal diode with a voltage drop of VF in series with a small internal resistance RD , as shown in Fig. 2-8(e). We can determine the approximate LED resistance RD from the slope of the linear section (above VF ) of the i–υ curve in Fig. 2-8(d); i.e., RD ≈ VD /I .

i = 0.5υ 2

2-1.5

RESISTIVE CIRCUITS

Kirchhoff’s Current Law (KCL)

As defined earlier, a node is a connection point for two or more branches. As such, it is not a real circuit element, and therefore it cannot generate, store, or consume electric charge. This assertion, which follows from the law of conservation of charge, forms the basis of Kirchhoff’s current law (KCL), which states that:  The algebraic sum of the currents entering a node must always be zero. 

“book” — 2015/5/4 — 6:57 — page 61 — #12

2-2

KIRCHHOFF’S LAWS

61 which states that:

i2 i 3 i1

 The total current entering a node must be equal to the total current leaving it. 

i4 How do we know which way a current is flowing in a circuit? Often, we do not. So, we guess by assigning a direction to each current, and then applying Kirchhoff’s laws to compute the currents. If the value for a particular current is a positive number, then our guess was correct, but if it is a negative number, then the direction of the current is opposite the one we assigned it.

Figure 2-9: Currents at a node.

Mathematically, KCL can be expressed by the compact form: N 

in = 0

(KCL),

(2.8)

n=1

Example 2-3: KCL Equations

Write the KCL equations at nodes 1 through 5 in the circuit of Fig. 2-10. Solution: At node 1: At node 2: At node 3: At node 4: At node 5:

where N is the total number of branches connected to the node, and in is the nth current. +” sign  A common convention is to assign a positive “+ −” to a current if it is entering the node and a negative “− sign if it is leaving it. 

−I1 − I3 + I5 = 0 I1 − I2 + 2 = 0 −2 − I4 + I6 = 0 −5 − I5 − I6 = 0 I3 + I4 + I2 + 5 = 0

For the node in Fig. 2-9, the sum of currents entering the node is

R1

(2.9)

 Alternatively, the sum of currents leaving a node is zero, in which case we assign a “+” to a current leaving the node and a “−” to a current entering it. 

V1 I1

I2 2A

R2

+_

where currents i1 and i4 were assigned positive signs because they are labeled in the figure as entering the node, and i2 and i3 were assigned negative signs because they are leaving the node.

+

i1 − i2 − i3 + i4 = 0,

2

R3

1

I3

I4

R4

5

3 I6

I5 R5

5A R6

Either convention is equally valid so long as it is applied consistently to all currents entering and leaving the node. By moving i2 and i3 to the right-hand side of Eq. (2.9), we obtain the alternative form of KCL, namely i1 + i4 = i2 + i3 ,

4 Figure 2-10: Circuit for Example 2-3.

(2.10)

“book” — 2015/5/4 — 6:57 — page 62 — #13

62

CHAPTER 2

Example 2-4: Applying KCL

 The algebraic sum of the voltages around a closed loop must always be zero. 

If V4 , the voltage across the 4  resistor in Fig. 2-11, is 8 V, determine I1 and I2 .



1 I1

N 

+_ 10 V −

+



I2

Figure 2-11: Circuit for Example 2-4.

υn = 0

(KVL),

(2.11)

n=1

+

_ V 4 2

This statement defines Kirchhoff’s voltage law (KVL). In equation form, KVL is given by





10 A

RESISTIVE CIRCUITS

where N is the total number of branches in the loop and υn is the nth voltage across the nth branch. Application of Eq. (2.11) requires the specification of a sign convention to use with it. Of those used in circuit analysis, the sign convention we chose to use in this book consists of two steps.

Sign Convention Solution: The designated direction of I2 is such that it enters the negative (−) terminal of V4 , whereas according to Ohm’s law, the current should enter through the positive (+) terminal of the voltage across a resistor. Hence, in the present case, we should include a negative sign in the relationship between I2 and V4 , namely I2 = −

V4 8 = − = −2 A. 4 4

Thus, the true direction of the current flowing through the 4  resistor is opposite of that of I2 . Using the KCL convention that defines a current as positive if it is leaving a node and negative if it is entering it, at node 2:

• Add up the voltages in a systematic clockwise movement around the loop. • Assign a positive sign to the voltage across an element if the (+) side of that voltage is encountered first, and assign a negative sign if the (−) side is encountered first. Hence, for the loop in Fig. 2-12, starting at the negative terminal of the 4 V voltage source, application of Eq. (2.11) yields −4 + V1 − V2 − 6 + V3 − V4 = 0.

(2.12)

10 − I1 + I2 = 0,

_ R1

V1 +

Kirchhoff’s Voltage Law (KVL)

The voltage across an element represents the amount of energy expended in moving positive charge from the negative terminal to the positive terminal, thereby establishing a potential energy difference between those terminals. The law of conservation of energy mandates that if we move electric charge around a closed loop, starting and ending at exactly the same location, the net gain or loss of energy must be zero. Since voltage is a surrogate for potential energy:

4V

-

R2

+

I1 = 10 + I2 = 10 − 2 = 8 A.

2-2.2

6V _

_ V + 2

which leads to

R3

+ _ + V4 _ R4 Figure 2-12: One-loop circuit.

+ V3 _

“book” — 2015/5/4 — 6:57 — page 63 — #14

2-2

KIRCHHOFF’S LAWS

63

Table 2-4: Equally valid, multiple statements of Kirchhoff’s Current Law (KCL) and Kirchhoff’s Voltage Law (KVL). ⎧ • ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎨• KCL ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ • ⎪ ⎪ ⎩

Sum of all currents entering a node = 0 [i = “+” if entering; i = “−” if leaving] Sum of all currents leaving a node = 0 [i = “+” if leaving; i = “−” if entering] Total of currents entering = Total of currents leaving

Solution: For the specified direction of I , we designate voltages V1 , V2 , and V3 across the three resistors, as shown in Fig. 2-13(b). In each case, the positive polarity of the voltage across a resistor is placed at the terminal at which the current enters the resistor. Starting at the negative terminal of the 12 V voltage source and moving clockwise around the loop, KVL gives −12 + V1 + V2 + V3 = 0. By Ohm’s law, V1 = 10I , V2 = 20I , and V3 = 30I . Hence,

⎧ • ⎪ ⎪ ⎪ ⎨

Sum of voltages around closed loop = 0 [υ = “+” if + side encountered first KVL ⎪ in clockwise direction] ⎪ ⎪ ⎩ • Total voltage rise = Total voltage drop

−12 + 10I + 20I + 30I = 0, which leads to 60I = 12, or I=

 An alternative statement of KVL is that the total voltage rise around a closed loop must equal the total voltage drop around the loop.  Recalling that a voltage rise is realized by moving from the (−) voltage terminal to the (+) terminal across the element, and voltage drop is the converse of that, the clockwise movement around the loop in Fig. 2-12 gives 4 + V2 + 6 + V4 = V1 + V3 ,

12 = 0.2 A. 60

I 20 Ω

10 Ω 12 V

+ _

30 Ω

(2.13)

(a) Circuit for Example 2-5 which mathematically conveys the same information contained in Eq. (2.12). Table 2-4 provides a summary of KCL and KVL statements.

I Concept Question 2-5: Explain why KCL is (in essence)

a statement of the law of conservation of charge. (See )

V1 = 10I

V2 = 20I

+

+

10 Ω 12 V

+ _

_

_

20 Ω 30 Ω

+ _V3 = 30I

Concept Question 2-6: Explain why KVL is a statement of conservation of energy. What sign convention is used with KVL? (See )

(b) After labeling voltages across resistors Example 2-5: Applying KVL

Determine the value of current I in the circuit of Fig. 2-13(a).

Figure 2-13: Circuit for Example 2-5 before and after labeling voltages across the three resistors with polarities consistent with Ohm’s law.

“book” — 2015/5/4 — 6:57 — page 64 — #15

64

R1

KCL/KVL Solution Recipe • Use KCL, KVL, and Ohm’s law to develop as many independent equations as the number of unknowns (N).

V0

(a) Write as many KVL loop equations as you can, picking up at least one additional circuit element for each loop. Let M be the number of such loop equations. Exclude loops that go through current sources.

• Cast the standard-form equations in matrix form, as in Eqs. (B.19) and (B.20) of Appendix B.

RESISTIVE CIRCUITS

R2

R4

+ _

R3

R5

I0

(a) Original circuit I1

(b) Write (N −M) KCL equations, making sure each node picks up an additional current. • Write the equations in standard form (see Eq. (B.2) in Appendix B).

CHAPTER 2

R1

R2

Va

I3

R4

I2

V0

+ _

I4 R5

R3 Loop 1

Vb

I0

Loop 2

• Apply matrix inversion to compute the values of the circuit unknowns (Appendix B).

(b) Example 2-6: Matrix Inversion of KVL/KCL Equations

For the circuit in Fig. 2-14(a): (a) identify all N unknown branch currents and assign them preliminary directions, (b) develop M KVL loop equations through all possible elements (while excluding loops containing current sources), (c) develop (N − M) KCL node equations, (d) arrange the equations in matrix form, (e) solve by matrix reduction to find the unknown branch currents, (f) determine the power dissipated in R5 , and (g) find the voltages of all extraordinary nodes relative to the negative terminal of the voltage source. The element values are: V0 = 10 V, I0 = 0.8 A, R1 = 2 , R2 = 3 , R3 = 5 , R4 = 10 , and R5 = 2.5 . Solution:

Figure 2-14: Circuit for Example 2-6.

(b) KVL equations The circuit contains two independent loops that do not contain the current source I0 . The associated KVL equations are: −V0 + I1 R1 + I1 R2 + R3 I2 = 0

(Loop 1),

−I2 R3 + I3 R4 + I4 R5 = 0

(Loop 2).

Alternatively, we can replace either of the two loop equations with the KVL equation for the perimeter loop that includes both of them, namely the loop that starts at the ground node, then goes clockwise through V0 , R1 , R2 , R4 , and R5 , and back to the ground node. Either approach leads to the same final result. (c) KCL equations

(a) Identify unknown currents Excluding the branch containing I0 (since we know that the current in that branch is I0 = 0.8 A), we have 4 unknown branch currents, which we denote I1 to I4 in Fig. 2-14(b). Also, with the negative terminal of the voltage source denoted as the voltage reference (ground), we have two extraordinary nodes, with designated voltages Va and Vb .

We have two extraordinary nodes (in addition to the ground node). We designate their voltages as shown in Fig. 2-14(b). With current defined as positive when entering a node, their KCL equations are

I1 − I2 − I3 = 0

(Node a),

I3 − I4 + I0 = 0

(Node b).

“book” — 2015/5/4 — 6:57 — page 65 — #16

2-2

KIRCHHOFF’S LAWS

65

(d) Arrange equations in matrix form





⎤ ⎤⎡ ⎤ ⎡ (R1 + R2 ) R3 0 0 V0 I1 ⎢ ⎢ ⎥ ⎢ ⎥ 0 −R3 R4 R5 ⎥ ⎢ ⎥ ⎢I2 ⎥ = ⎢ 0 ⎥ . ⎣ ⎣ ⎣ ⎦ ⎦ 0 ⎦ 1 −1 −1 0 I3 −I0 0 0 1 −1 I4      

A

I



+ _

Vab

+ _

12 V

This is in the form

+



B



a

24 V

_

AI = B.

b (a) Given circuit

(e) Matrix inversion

6Ω After replacing the sources and resistors with their specified numerical values, matrix reduction, per MATLAB, MathScript, or the procedure outlined in Appendix B-2, leads to I1 = 1.1 A,

I4 = 1 A, .

(f) Power in R5 P =

I3 I2

I2 = 0.9 A,

I3 = 0.2 A,

Node 1 I1



_ V1 = 2I2

Vb = I4 R5 = 1 × 2.5 = 2.5 V.



_ 6V +

1A 1A 2Ω

Example 2-7: Two-Source Circuit

Solution: The circuit contains two independent loops and two extraordinary nodes, which we label node 1 and node 2 in Fig. 2-15(b). At extraordinary node 1, we assign currents I1 , I2 , and I3 . Their directions are chosen arbitrarily; for I1 , for example, if the solution yields a positive value, then the direction we assigned it is indeed the correct direction, and if the solution yields a negative value, then its true direction is the opposite of what we had assigned it. Once the directions of I1 to I3 are specified at node 1, continuity of current automatically specifies their directions at node 2, as shown in Fig. 2-15(b). Since we have 3 unknowns (I1 , I2 , and I3 ), we need N = 3 equations.

24 V

b (b) After assigning currents at nodes 1 and 2

Va = I2 R3 = 0.9 × 5 = 4.5 V,

Determine Vab in the circuit of Fig. 2-15(a).

I1 + _

L1

+ _

12 V

Node 2

V2 = 4I2

V4 = 3I1

2

(g) Node voltages

I2

_

+

+

= 1 × 2.5 = 2.5 W.

I3 4Ω

a

_

+

3Ω I42 R5

_

+

V3 = 6I3 L2

Node 1 2A

_ + _

6V

3Ω 12 V

+ _

2V+



a

+

_

Node 2 4V

+

Vab

+ _

_ b (c) After completing solution

Figure 2-15: Circuit for Example 2-7.

24 V

“book” — 2015/5/4 — 6:57 — page 66 — #17

66

CHAPTER 2

In terms of the labeled voltages, application of KVL around the two loops gives −12 + V4 + V1 + V2 + 24 = 0,

(KVL for Loop 1) (2.14a)

V3 − V2 − V1 = 0.

(KVL for Loop 2) (2.14b)

Using Ohm’s law for the four resistors, the two KVL equations become −12 + 3I1 + 2I2 + 4I2 + 24 = 0, 6I3 − 4I2 − 2I2 = 0.

(KVL for Loop 1) (2.15a) (KVL for Loop 2) (2.15b)

The two simultaneous equations contain three unknowns, namely I1 to I3 . A third equation is supplied by KCL at node 1 or node 2: I1 = I2 + I3 .

(KCL @ node 1 or 2)

(2.16)

voltage rise of 6 V, and from node 1 to node b is a third voltage rise of 2 V. Hence Vab = 12 + 6 + 2 = 20 V. Alternatively, we can calculate Vab by moving from node b to node a counterclockwise through node 2. In that case Vab = 24 − 4 = 20 V, which is identical to the earlier result. Example 2-8: Circuit with Dependent Source

The circuit in Fig. 2-16 includes a current-dependent voltage source. Apply KVL and KCL to determine the amount of power consumed by the 12  resistor. Solution: We start by assigning currents I2 and I3 at node 1, and using those currents to designate the voltages across the three resistors. Note that in all cases, the designated (+) side of

Equations (2.15a), (2.15b), and (2.16) constitute 3 equations in 3 unknowns. We can solve for I1 to I3 either by the substitution method or by matrix inversion (Appendix B). To apply the latter, we need to cast the three equations in standard form: 3I1 + 6I2

= −12,

RESISTIVE CIRCUITS

I1



20 V

12 Ω

+ _



_ +

8I1

_ +

8I1

−6I2 + 6I3 = 0, I1 − I2 − I3 = 0.

(a) Given circuit

In matrix form: ⎡

⎤⎡ ⎤ ⎡ ⎤ 3 6 0 I1 −12 ⎣0 −6 6 ⎦ ⎣I2 ⎦ = ⎣ 0 ⎦ . 1 −1 −1 0 I3

+

Matrix inversion, as outlined in Appendix B, leads to I1 = −2 A,

I2 = −1 A,

I3 = −1 A.

Hence, the true directions of the three currents are exactly opposite those we supposed , and so are the polarities of the voltages across the resistors. Incorporating both the calculated magnitudes and signs of I1 to I3 leads to the diagram shown in Fig. 2-15(c). To calculate Vab , we start at node b and move clockwise towards node a in loop 1, while keeping track of voltage rises and drops. From node b to the (+) terminal of the 12 V source is a voltage rise of 12 V, from there to node 1 is a

I1



20 V

Node 1 I3

_ 4I1

I2

+ 8I2 _

+ _

12 Ω

+

_ 12I3



L1

L2

Node 2 (b) After assigning currents at node 1 Figure 2-16: Circuit for Example 2-8.

“book” — 2015/5/4 — 6:57 — page 67 — #18

2-3

EQUIVALENT CIRCUITS

67

the voltage corresponds to the terminal at which the current is entering. For loops 1 and 2, KVL gives −20 + 4I1 + 8I2 = 0,

(KVL for Loop 1)

−8I2 + 12I3 − 8I1 = 0.

(KVL for Loop 2)

Note that there is another loop in the circuit, namely the perimeter loop around the whole circuit, but if we write a KVL equation for that loop, it would not provide an equation independent of the other loop equations because it would not include any circuit element not already included in loops L1 and L2 . At node 1, KCL states that I1 = I2 + I3 .

Exercise 2-5: Apply KCL and KVL to find I1 and I2 in

Fig. E2.5.

4 Ω I2 I1 + _

20 V



4A

Figure E2.5 Answer: I1 = 6 A, I2 = 2 A. (See

)

Exercise 2-6: Determine Ix in the circuit of Fig. E2.6.

The combination of the three equations in unknowns I1 , I2 , and I3 leads to the solution

4A 2Ω

25 A, I1 = 7 5 I2 = A, 7 20 I3 = A. 7

Ix 4Ω

2Ω _ +



2Ix

Figure E2.6

Hence, the power dissipated in the 12  resistor is  P = I32 R =

20 7

Answer: Ix = 1.33 A. (See

2 × 12 = 97.96 W.

2-3 Exercise 2-4: If I1 = 3 A in Fig. E2.4, what is I2 ?



I2 I1

10 V

+ _



2A

Figure E2.4 Answer: I2 = −1 A. (See

)

)

Equivalent Circuits

Even though Kirchhoff’s current and voltage laws can be used to write down the requisite number of node and loop equations that are necessary to solve for all of the voltages and currents in a circuit, it is often easier to determine a certain unknown voltage or current by first simplifying the other parts of the circuit. The simplification process involves the use of circuit equivalence, wherein a circuit segment connected between two nodes (such as the original circuit segment connected between nodes 1 and 2 in Fig. 2-17) is replaced with another, simpler, circuit whose behavior is such that the voltage difference (υ1 − υ2 ) between the two nodes—as well as the currents entering into them (or exiting from them)—remain unchanged. That is:

“book” — 2015/5/4 — 6:57 — page 68 — #19

68

CHAPTER 2

Circuit Equivalence Original circuit segment

i1

υ1

i2

1 υ2

RESISTIVE CIRCUITS

Combining In-Series Resistors is Rest of the circuit

υs

1

+ + __

2 2

R1

R2

υ1

υ2

υ5

υ4

R5

R4

υ3

R3

(a) Original circuit

Equivalent circuit

i1

υ1

i2

1 υ2

is

Rest of the circuit υs

2

1

+ + __

Equivalent circuit

Figure 2-17: Circuit equivalence requires that the equivalent circuit exhibit the same i–υ characteristic as the original circuit.

Req

2 (b) Req = R1 + R2 + R3 + R4 + R5

 Two circuits connected between a pair of nodes are considered to be equivalent—as seen by the rest of the circuit—if they exhibit identical i–υ characteristics at those nodes.  To the rest of the circuit, the original and equivalent circuit segments appear identical. The equivalent-circuit technique can be applied on the source side of a circuit, as well as on the load side. We now will examine several types of equivalent circuits and then provide an overall summary at the conclusion of this section.

2-3.1

Resistors in Series

Consider the single-loop circuit of Fig. 2-18(a) in which a voltage source υs is connected in series with five resistors. The KVL equation is given by −υs + R1 is + R2 is + R3 is + R4 is + R5 is = 0,

(2.17)

which can be rewritten as

(2.18)

where Req is an equivalent resistor whose resistance is equal to the sum of the five in-series resistances, Req = R1 + R2 + R3 + R4 + R5 .

From the standpoint of the source voltage υs and the current is it supplies, the circuit in Fig. 2-18(a) is equivalent to that in Fig. 2-18(b). That is, υs is = . (2.20) Req  Multiple resistors connected in series (experiencing the same current) can be combined into a single equivalent resistor Req whose resistance is equal to the sum of all of their individual resistances.  Mathematically,

Req =

N 

Ri

(resistors in series),

(2.21)

i=1

where N is the total number of resistors in the group.

υs = R1 is + R2 is + R3 is + R4 is + R5 is = (R1 + R2 + R3 + R4 + R5 )is = Req is ,

Figure 2-18: In a single-loop circuit, Req is equal to the sum of the resistors.

(2.19)

Voltage division For resistor R2 in Fig. 2-18(a), the voltage across it is given by   R2 υ2 = R2 is = υs . (2.22a) Req

“book” — 2015/5/4 — 6:57 — page 69 — #20

2-3

EQUIVALENT CIRCUITS

69

Similar expressions apply to the other resistors, wherein the voltage across a resistor is equal to υs multiplied by the ratio of its own resistance to the sum total Req . Thus, the single-loop circuit, in effect, divides the source voltage among the series resistors.  The voltage across any individual resistor Ri in a series circuit is a proportionate fraction (Ri /Req ) of the voltage across the entire group  υi =

Ri Req

 υs

(voltage division).

(2.22b) 

Example 2-9: The Voltage Divider

supply a secondary load circuit a specific voltage υ2 that is smaller than the available source voltage υs . In other words, the goal is to scale υs down to υ2 . If υs = 100 V, choose appropriate values for R1 and R2 so that υ2 = 60 V. Solution: In view of Eq. (2.22a), application of the voltagedivision property gives  υ2 =

R2 υ2 60 = = = 0.6, R1 + R 2 υs 100

 (a)

υ2 =

R2 R1 + R 2

R1 = 2 

and

R2 = 3 .

Note that the circuit in Fig. 2-19(b) will provide approximately the indicated voltages to a load circuit, so long as the resistance of the load circuit is very large compared with the resistance of R2 . Otherwise, the load circuit would draw current, thereby “loading down” the source circuit and changing V2 .

+ υ2 _

R2

υs .

which can be satisfied through an infinite combination of choices of R1 and R2 . Hence, we arbitrarily choose

R1 υs



To obtain the desired division, we require

The term voltage divider is used commonly in reference to a circuit of the type shown in Fig. 2-19, whose purpose is to

+ _

R2 R1 + R 2

Load circuit

2-3.2

 υs

+

+

Sources in Series

Figure 2-20 contains a single-loop circuit composed of a voltage source, a resistor, and two current sources, all connected in series. One of the current sources indicates that the current flowing through it is 4 A in magnitude and clockwise in direction, while the other current source indicates that the

6V 3Ω

+ 6V _



+ 4V _

_

_

10 V +

R

+

6A

4V _

_

V0

+ _

4A

(b) Voltage divider is equivalent to subdividing a battery into two separate batteries Figure 2-19: Voltage dividers are important tools in circuit analysis and design.

Figure 2-20: Unrealizable circuit; two current sources with different magnitudes or directions cannot be connected in series.

“book” — 2015/5/4 — 6:57 — page 70 — #21

70

TECHNOLOGY BRIEF 4: RESISTIVE SENSORS

Technology Brief 4 Resistive Sensors Resistive sensors can convert many physical parameters in our environment into a resistance that varies with temperature, light, pressure, moisture, chemical composition, sound, or other inputs. This variable resistance will then change the voltage or current in a circuit, which can be further manipulated in an electrical system to produce a desired output (turning on a warning light or buzzer, adjusting a valve, or otherwise control the pressure/light/heat/sound automatically). When a system measures a parameter (e.g., temperature) in order to control that parameter, the process is called a feedback loop. Sensors are a very important part of a feedback system. So how do resistive sensors work? The resistance R of a semiconductor accounts for the reduction in the electrons’ velocities due to collisions with the much larger atoms of the conducting material (see Technology Brief 3). The question is:What happens to R if we disturb the atoms of the conductor by applying an external, nonelectrical stimulus, such as heating or cooling it, stretching or compressing it, or shining light on it? Through proper choice of materials, we can modulate (change) the magnitude of R in response to such external stimuli.

Piezoresistive Sensors (Pressure, Bending, Force, etc.) In 1856, Lord Kelvin discovered that applying a mechanical load on a bar of metal changed its resistance. Over the next 150 years, both theoretical and practical advances made it possible to describe the physics behind this effect in both conductors and semiconductors. The phenomenon is referred to as the piezoresistive effect (Fig. TF4-1) and is used in many practical devices to convert a mechanical signal into an electrical one. Such sensors (Fig. TF4-2) are called strain gauges. Piezoresistive sensors are used in a wide variety of consumer applications, including writing styluses for tablets (some high-precision styluses are resistive and others are capacitive—which we will learn about in Chapter 5), robot toy “skins” that sense force, microscale gas-pressure sensors, and micromachined accelerometers that sense acceleration. They all use piezoresistors in electrical circuits to generate a signal from a mechanical stimulus.

R (Ω)

STRETCHING F

F

F

F

COMPRESSION

FORCE (N)

F=0

R=ρ

l A

Figure TF4-1: Piezoresistance varies with applied force. The word “piezein” means “to press” in Greek.

In its simplest form, a resistance change R occurs when a mechanical pressure P (N/m2 ) is applied along the axis of the resistor (Fig. TF4-1) R = R0 αP, where R0 is the unstressed resistance and α is known as the piezoresistive coefficient (m2 /N).The piezoresistive coefficient is a material property, and for crystalline materials (such as silicon), the piezoresistive coefficient also varies depending on the direction of the applied pressure (relative to the crystal planes of the material). When the horizontal and vertical components are different the material is called anisotropic. The total resistance of a piezoresistor under stress is therefore given by R = R0 + R = R0 (1 + αP). The pressure P, which usually is called the mechanical stress or mechanical load, is equal to F/A, where F is the force acting on the piezoresistor and A is the cross-sectional area it is acting on. The sign of P is defined as positive for a compressive force and negative for a stretching force. The piezoresistive coefficient α usually has a negative value, so the product αP leads to a decrease in R for compression and an increase for stretching.

Thermistor Sensors Changes in temperature also can lead to changes in the resistance of a piece of conductor or semiconductor;

“book” — 2015/5/4 — 6:57 — page 71 — #22

TECHNOLOGY BRIEF 4: RESISTIVE SENSORS

71

(a)

(b)

(c)

Figure TF4-2: A microfabricated pressure sensor utilizing piezoresistors as sensors. (a) A thin diaphragm (blue) is suspended over a depression etched into a glass substrate (grey). Serpentine piezoresistors (yellow) are patterned onto the membrane. (b) Differences in pressure between between the ambient and the gas in the depression will move the membrane. When this happens, the resistors stretch (or compress), changing their resistance as explained in the text. (c) A false color scanning electron micrograph of an actual microfabricated pressure sensor. Note the piezoresistors (yellow) patterned along the four sides of the diaphragm and the white, 100 μm scale bar. (Courtesy of Khalil Najafi, University of Michigan.) when used as a sensor, such an element is called a thermistor. As a simple approximation, the change in resistance can be modeled as R = k T, where T is the temperature change (in degrees C) and k is the first-order temperature coefficient of resistance (/◦ C). Thermistors are classified according to whether k is negative or positive (i.e., if an increase in temperature decreases or increases the resistance). This approximation works only for small temperature changes; for larger swings, higher-order terms must be included in the equation. Resistors used in electrical circuits that are not intended to be used as sensors are manufactured from materials with the lowest k possible, since circuit designers do not want their resistors changing during operation. In contrast, materials with high values of k are desirable for sensing temperature variations. Care must be taken, however, to incorporate into the sensor response the self-heating effect that occurs due to having a current passing through the resistor itself (as in the flow sensor shown in Fig. TF4-3).

Thermistors are used routinely in modern thermostats, cell phones, automotive and industrial applications, weather monitoring, and battery-pack chargers (to prevent batteries from overheating). Thermistors also have found niche applications (Fig. TF4-3) in lowtemperature sensing and as fuse replacements (for thermistors with large, positive k values). In the case of current-limiting fuse replacements, a large enough current self-heats the thermistor, and the resistance increases. There is a threshold current above which the thermistor cannot be cooled off by its environment; as it continues to get hotter, the resistance continues to increase, which in turn, causes even more selfheating. This “runaway” effect rapidly shuts current off almost entirely. Thermistors are specified based on their linear range where resistance varies linearly with the temperature, and a wide variety of options are available.

Moisture and Chemical Sensors Resistive sensors can also be built with two electrodes measuring the material between them. A simple moisture

“book” — 2015/5/4 — 6:57 — page 72 — #23

72

TECHNOLOGY BRIEF 4: RESISTIVE SENSORS

Figure TF4-3: This micromachined anemometer (flow meter) is a thermistor that measures fluid velocity. The resistor (red) serves as both a heater and a thermistor. During operation, a voltage across the resistor produces a current (I = V /R) which heats the resistor (recall the heat power, P = V ∗ I ). As fluid flows by the resistor (blue), the flow draws away heat. Since increasing the flow increases the cooling of the resistor and temperature changes the resistance, the flow can be inferred from the thermistor. (Courtesy of Khalil Najafi, University of Michigan.)

sensor you can build yourself consists of two electrodes with an absorbing material between them (Fig. TF4-4). Just draw two thick pencil (graphite) lines on paper, clip to them with alligator clips, and measure the resistance with your myDAQ. Then drip water between the two lines, so that it makes contact between them. The resistance will immediately drop in magnitude. In a similar approach, resistive sensors can sometimes be used to determine chemical composition of a liquid

material. The resistivity of the material depends strongly on the number of dissolved or loose ions in the material (see Table 2-1). Deionized water has high resistivity, drinking water has moderate resistivity, and sea water has low resistivity. Placing two electrodes into a container of fluid, or running fluid over two electrodes in a microfluidic system can be used to measure the resistivity of the material and hence its chemical composition.This is often used as a simple way to monitor the purity of drinking water.

Graphite

Drip water

Alligator clips

Figure TF4-4: Increased ions (from dissolved solids, for example) increase the conductivity (reduce resistivity), which can be measured by an ohmmeter.

“book” — 2015/5/4 — 6:57 — page 73 — #24

2-3

EQUIVALENT CIRCUITS

73

current is 6 A in magnitude and counterclockwise in direction. Continuity of current flow mandates that the current flowing through the loop be exactly the same in both magnitude and direction at every location over the full extent of the loop. So our dilemma is: Is the current 4 A, 6 A, or the difference between the two? It is none of those guesses. The true answer is that the circuit is unrealizable, meaning that it is not possible to construct a circuit with two current sources of different magnitudes or different directions that are connected in series. The problem with the circuit of Fig. 2-20 has to do with our representation of ideal current sources. As was stated in Section 1-6.2 and described in Table 1-5, a real current source can be modeled as the parallel combination of an ideal current source and a shunt resistor Rs . Usually, Rs is very large, so very little current flows through it in comparison with the current flowing through the other part of the circuit, in which case it can be deleted without much consequence. In the present case, however, had such shunt resistors been included in the circuit of Fig. 2-20, the dilemma would not have arisen. The lesson we should learn from this discussion is that when we idealize current sources by deleting their parallel resistors, we should never connect them in series in circuit diagrams.

υ2 +_

R1

R2

-

+ _

υ1

_ + Node 1

υ3

Node 2 RL

(a)

υeq

+ _

Req Node 1

Node 2 RL

 Ideal current sources cannot be added in series. 

(b) υeq = υ1 − υ2 + υ3

Whereas current sources cannot be connected in series, voltage sources can. In fact, it follows from KVL that from the standpoint of an external load resistor RL connected between nodes 1 and 2, the circuit in Fig. 2-21(a) can be simplified into the equivalent circuit of Fig. 2-21(b) with

υeq = υ1 − υ2 + υ3

(2.23)

Req = R1 + R2 .

(2.24)

and

Thus:

Figure 2-21: In-series voltage sources can be added together algebraically.

2-3.3

Resistors and Sources in Parallel

When multiple resistors are connected in series, they all share the same current, but each has its own individual voltage across it. The converse is true for multiple resistors connected in parallel: the three resistors in Fig. 2-22(a) experience the same voltage across all of them, namely υs , but each carries its own individual current. The current supplied by the source is divided among the branches containing the three resistors. Thus, is = i1 + i2 + i3 .

(2.25)

Application of Ohm’s law provides i1 =

 Multiple voltage sources connected in series can be combined into an equivalent voltage source whose voltage is equal to the algebraic sum of the voltages of the individual sources. 

Req = R1 + R2

υs , R1

i2 =

υs , R2

and

i3 =

υs , R3

(2.26)

which when used in Eq. (2.25) leads to is =

υs υs υs + + . R1 R2 R3

(2.27)

“book” — 2015/5/4 — 6:57 — page 74 — #25

74

CHAPTER 2

1

υs

This result can be generalized to any N resistors connected in parallel

is

+ _

i1

i2

i3

R1

R2

R3

N

 1 1 = Req Ri

(resistors in parallel).

Current division

(a) Original circuit

 Multiple resistors connected in parallel divide the input current among them.  For R2 in Fig. 2-22(a),

is

υs = R2

i2 =

υs

+ _

Req

(b) 

1 1 1 + + R1 R2 R3

−1

 i2 =

Req R2



Req R2

 is .

(2.32)

By extension, for a current divider composed of N in-parallel resistors, the current flowing through Ri is a proportionate fraction (Req /Ri ) of the input current. It is useful to note that the equivalent resistance for a parallel combination of two resistors R1 and R2 (Fig. 2-23) is given by

2

Req =

(2.31)

i=1

2

1

RESISTIVE CIRCUITS

 is

Req =

R 1 R2 . R1 + R 2

(2.33)

Figure 2-22: Voltage source connected to a parallel combination of three resistors.

 As a short-hand notation, we will sometimes denote such a parallel combination R1  R2 . We also denote the series combination of R1 and R2 as (R1 + R2 ). 

We wish to replace the parallel combination of the three resistors with a single equivalent resistor Req , as depicted in Fig. 2-22(b), such that the current is remains unchanged. For the equivalent circuit, υs . (2.28) is = Req If the two circuits in Fig. 2-22 are to function the same, as regards the source, then is as given by Eq. (2.27) for the original circuit should be equal to the expression for is given by Eq. (2.28) for the equivalent circuit. Thus, υs υs υs υs = + + , Req R1 R2 R3

(2.29)

Current Division 1

is

(2.30)

1 i2

i1 R1

R1R2 Req = R + R 1 2

R2

2

2  i1 =

from which we conclude that 1 1 1 1 = + + . Req R1 R2 R3

As was noted earlier in Section 2-1.5, the inverse of the resistance R is the conductance G; G = 1/R. For N conductances

R2 R1 + R 2



 is

i2 =

R1 R1 + R 2

 is

Figure 2-23: Equivalent circuit for two resistors in parallel.

“book” — 2015/5/4 — 6:57 — page 75 — #26

2-3

EQUIVALENT CIRCUITS

75 can be combined when connected in series, but they cannot be connected in parallel, unless they have identical voltages (Fig. 2-24). Two current sources can be combined when connected in parallel (as illustrated by Fig. 2-25), but they cannot be connected in series.

+

+

+

V1

V2

V3

_

_

Example 2-10: Current Division Using Conductance

_

For the circuit in Fig. 2-26: (a) Relate I3 to I0 and resistances R1 to R3 . Figure 2-24: This is an unrealizable circuit unless all voltage sources have identical voltages and polarities; that is, V1 = V2 = V3 .

connected in parallel, Eq. (2.31) assumes the form of a linear sum

Geq =

N 

(b) Relate I3 to I0 and conductances G1 to G3 , where Gi = 1/Ri . Solution: (a) Application of the expressions given in Fig. 2-22 leads to   Req I3 = I0 , R3 with

Gi

(conductances in parallel).

(2.34)

i=1

Two resistors always can be combined together, whether they are connected in series (sharing the same current) or in parallel (sharing the same voltage). Two voltage sources

 Req =

1 1 1 + + R1 R2 R3

I3 =

3 I1

R2

I2

R3

I3

2 Node 2

1

Req

R1 R2 R3 R 2 R3 + R 1 R3 + R 1 R2

Req R3



 I0 =

R 1 R2 R 2 R3 + R 1 R3 + R 1 R2

with 1 = Req



1 1 1 + + R1 R2 R3

Req

.

 I0 .

= G1 + G2 + G3 .

Ieq

I3 I0

R1

R2

Ieq = I1 − I2 + I3

Figure 2-25: Adding current sources connected in parallel.





2 R2 R 3 = R2  R3 = R2 + R 3

−1

(b) Rewriting the expressions for I3 and Req in terms of conductances gives   G3 I3 = I0, Geq

Geq =

R1



R2 R3 + R1 R3 + R1 R2 R1 R2 R3

Hence,

Node 3

R1

 = =



1

−1

Figure 2-26: Circuit of Example 2-10.

R3

“book” — 2015/5/4 — 6:57 — page 76 — #27

76

CHAPTER 2

Hence,

 I3 =

G3 G1 + G 2 + G 3

RESISTIVE CIRCUITS

(b) Circuit of Fig. 2-27(b): Circuit is realizable. From the standpoint of the two voltage sources to the left of nodes CD,

 I0 .

Current division using conductances assumes the same functional form as voltage division using resistances (Eq. (2.22b)). Example 2-11: Realizable and Unrealizable Circuits

Given that the voltage difference between any two nodes in a circuit has to be unique (cannot have multiple values simultaneously), and that the current in any given branch also is unique, determine which of the three circuits in Fig. 2-27 are realizable and which are unrealizable. Solution: (a) Circuit of Fig. 2-27(a): Circuit is not realizable. From the perspective of the ideal voltage source Vs , the voltage difference between nodes A and B is Vs , but according to the dependent source the voltage is 2Vs .

VCD = V1 + V2 = 20 − 5 = 15 V. Also connected across nodes CD is voltage source V3 , but its voltage is exactly 15 V. Two voltage sources can be connected in parallel if they have the same voltage. (c) Circuit of Fig. 2-27(c): Circuit is realizable. KCL at node E requires that the sum of the three currents entering the node be zero. Hence, 3 + 2Ix − Ix = 0, which leads to Ix = −3 A. This means that the direction of Ix is upwards and the dependent current source has a downward-moving current of 6 A. Example 2-12: Equivalent-Circuit Solution

A Vs

+ _

10 Ω

+ _

2Vs

20 Ω V3

+ _

15 V

B (a) V2 +_ V1

+ _

C

20 V 5 V

Use the equivalent-resistance approach to determine V2 , I1 , I2 , and I3 in the circuit of Fig. 2-28(a). Solution: In the circuit of Fig. 2-28(a), the part of the circuit connected to the voltage source is equivalent to a resistor Req = R1 + [(R3 + R4 )  (R2 + R5 )]. Hence, our first step is to combine the 2  and 4  in-series resistances into a 6  resistance and to combine the two 6  in-parallel resistances into a 3  resistance (by applying Eq. (2.33)). The simplifications lead to the circuit in Fig. 2-28(b). Next, we calculate the parallel combination of the 3  and 6  resistors, (3  6), again using Eq. (2.33), to get (3 × 6)/(3 + 6) = 18/9 = 2 . The new equivalent circuit is displayed in Fig. 2-28(c), from which we deduce that

D

I1 =

(b) and

E

V2 = 2I1 = 2 × 2 = 4 V.

Ix 3A

24 = 2A 10 + 2

30 Ω (c) Figure 2-27: Circuits of Example 2-11.

2Ix

Returning to Fig. 2-28(b), we apply Ohm’s law to find I2 and I3 . I2 =

V2 4 = = 1.33 A, 3 3

I3 =

V2 4 = = 0.67 A. 6 6

and

“book” — 2015/5/4 — 6:57 — page 77 — #28

2-3

EQUIVALENT CIRCUITS

V1 10 Ω 1

I1

I2

R2

+

24 V

+ -_

R3

Concept Question 2-8: What is a voltage divider and



V2 2

R1

77

R4





I3 R5



(a)

1

I1

a conductance G? (See

V2 I2



I3

10 V



+ _

2

Combining 3 Ω and 6 Ω in parallel

1

R1

I1

V2 2

+

+ -_



1

1

1

1

2 2

1

1

1

1

Figure E2.7 Answer: I = 5 A. (See

(b)

V1 10 Ω

)

I

2

+ 24 V + -_

24 V

Concept Question 2-9: What is the i–υ relationship for

1 2

R1

)

Exercise 2-7: Apply resistance combining to simplify the circuit of Fig. E2.7 in order to find I . All resistor values are in ohms.

Combine R3 and R4 in parallel

V1 10 Ω

what is a current divider? (See

2-3.4

)

Source Transformation

We now will demonstrate how a realistic voltage source composed of an ideal voltage source in series with a resistor can be exchanged for a realistic current source composed of an ideal current source in parallel with a shunt resistor, or vice versa. The two circuits are shown in parts (a) and (b) of Fig. 2-29. Exchanging the one source for the other requires that they be equivalent—from the vantage point of the external circuit.

(c)

 A voltage-source circuit and a current-source circuit are considered equivalent and interchangeable if they deliver the same input current i and voltage υ12 to the external circuit. 

Figure 2-28: Example 2-12. (a) Original circuit, (b) after combining R3 and R4 in parallel and combining R2 and R5 in series, and (c) after combining the 3  and 6  resistances in parallel.

For the voltage-source circuit, application of KVL gives −υs + iR1 + υ12 = 0,

(2.35)

from which we obtain the following expression for i: Concept Question 2-7: What conditions must be satisfied

in order for two circuits to be considered equivalent? (See )

i=

υs υ12 − . R1 R1

(2.36)

“book” — 2015/5/4 — 6:57 — page 78 — #29

78

CHAPTER 2 In summary:

Source Transformation i

R1

υs

1

+

+ -_

 A voltage source υs in series with a source resistance Rs is equivalent to the combination of a current source is = υs /Rs , in parallel with a shunt resistance Rs . The direction of the equivalent current source is the same as the direction from (−) to (+) terminals of the voltage source. 

υ12 External circuit

678

2

Voltage source (a) is

i

This equivalence is called source transformation because it allows us to replace a realistic voltage source with a realistic current source, or vice versa. A summary of in-series and in-parallel equivalent circuits involving sources and resistors is available in Table 2-5.

1 External circuit

iR2 is

RESISTIVE CIRCUITS

R2

υ12

Example 2-13: Source Transformation

Determine the current I in the circuit of Fig. 2-30(a).

678

2

Current source is = υs /R1 R2 = R1

Solution: It is best to avoid transformations that would involve the 3  resistor with the unknown current I . Hence, we will apply multiple source-transformation steps, moving from the left end of the circuit towards the 3  resistor.

(b) Figure 2-29: Realistic voltage and current sources connected to an external circuit. Equivalence requires that is = υs /R1 and R2 = R1 .

Step 1: Current to voltage transformation allows us to convert the combination (Is1 , Rs1 ) to a voltage source Vs1 = Is1 Rs1 = 16 × 2 = 32 V, in series with Rs1 .

Application of KCL to the current-source circuit gives i = is − iR2

υ12 = is − , R2

Step 2: Combining Rs1 in series with the 6  resistor results in (2.37) Rs2 = 2 + 6 = 8 .

where we used Ohm’s law to relate iR2 to υ12 . Equivalence of Eqs. (2.36) and (2.37) is satisfied for all values of i and υ12 if and only if:

Hence, the new input source becomes (Vs1 , Rs2 ). Step 3: Convert (Vs1 , Rs2 ) back into a current source

R1 = R2

(2.38a)

υs . R1

(2.38b)

Is2 = Vs1 /Rs2 = 32/8 = 4 A,

and is =

in parallel with Rs2 . Step 4: Combine Rs2 = 8  in parallel with the other 8  resistor (8  8) to obtain an equivalent resistance Rs3 = 4 .

“book” — 2015/5/4 — 6:57 — page 79 — #30

2-3

EQUIVALENT CIRCUITS

79

Table 2-5: Equivalent circuits.

Circuit

Equivalent 1

R1

Series

R1

R1 + R2

R2

2 c

R2

Y

R3 3

R1

Parallel

R2 1 R1 1 G2 = R2 G1 =

+ _

Series

+ _

(R1 || R2)

R1R2 R1 + R2

(G1 || G2)

G1 + G 2

Rb

υ1

+ _

Ra

R2 = i2

i1 + i2

R3 = Ra = Rb =

Rs is =

υs

Source transformation

Step 5: Convert again to a voltage source

υs Rs

Rc =

Rs

Rb Rc Ra + Rb + Rc Ra Rc Ra + Rb + Rc Ra Rb Ra + Rb + Rc R1 R2 + R2 R3 + R1 R3 R1 R1 R2 + R2 R3 + R1 R3 R2 R1 R2 + R2 R3 + R1 R3 R3

For Ra = Rb = Rc For R1 = R2 = R3

R1 = R2 = R3 = Ra / 3 Ra = Rb = Rc = 3R1

For the single loop realized in the final step,

Vs2 = Is2 Rs3 = 4 × 4 = 16 V, in series with Rs3 .



3

υ2

Parallel i1

2

υ1 + υ2

R1 =

+ _

Rc

1

I=

16 Vs2 = = 2 A. 4+1+3 8

“book” — 2015/5/4 — 6:57 — page 80 — #31

80

CHAPTER 2



Solution:

1Ω I=?

Is1 = 16 A

Rs1 = 2 Ω

Step 2 (series R) + Vs1 = 32 V + -_

Rs = 8 Ω 678



RESISTIVE CIRCUITS

Step 1: Convert the 2 A current source in parallel with the 20  resistor into a 40 V voltage source in series with a 20  resistor.



Step 2: Combine the two in-series 20  and 40  resistances into a 60  resistance, and combine the 40 V and 16 V sources into a single 24 V source.

Step 1 (source transformation)

Step 3: Convert each voltage source (together with its in-series resistance) into a current source with a resistance in parallel.

2







I

Rs1 8Ω

Step 4: Combine the two in-parallel resistances and the two in-parallel current sources. Step 5: For RL = 10 , current division yields

3Ω I=

Step 3 (source transformation) 1Ω

20 × 3 = 2 A, 10 + 20

and the associated voltage across RL is Vab = 10I = 20 V.

I I s2 = 4 A

Rs2 = 8 Ω





Exercise 2-8: Apply source transformation to the circuit

in Fig. E2.8 to find I . Answer: I = 4 A. (See

Step 4 (parallel R)

6Ω I

1Ω I Is2 = 4 A

Rs3 = 4 Ω

)

12 V

+ _





10 A



Figure E2.8

4Ω Vs2 = 16 V

Step 5 (source transformation) 1Ω

+

+ -_

I 3Ω

Figure 2-30: Example 2-13 circuit evolution.

Example 2-14: Finding Vab

While keeping the load resistor RL in the top circuit of Fig. 2-31 intact, apply source transformations until the circuit simplifies to a current divider, then determine Vab for RL = 10 .

2-4 Wye–Delta (Y–) Transformation In principle, it always is possible to simplify the behavior of a resistive circuit when measured across any two nodes—no matter how complex its topology—down to a simple equivalent circuit composed of an equivalent voltage source in series with an equivalent resistor. The preceding sections offered us tools for combining resistors together whenever they are connected in series or in parallel, as well as for combining in-series voltage sources and in-parallel current sources. Sometimes, however, we may encounter circuit topologies that cannot be simplified using those tools because their resistors are connected neither in series nor in parallel. A case in point is the circuit in Fig. 2-32, in which no two resistors share the same current or voltage. This

“book” — 2015/5/4 — 6:57 — page 81 — #32

2-4 WYE–DELTA (Y–) TRANSFORMATION

102 V +_

+ _

16 V

30 Ω RL

a

81

b

3A

20 Ω

40 Ω

2A 20 Ω

Step 1 (source transformation) 102 V +_

30 Ω

I

Step 4 (parallel I and R)

40 V 30 Ω

Step 2 (series V and R)

24 V

30 Ω

a

RL

b

60 Ω

RL _ +

b

3.4 A

+ _

40 Ω

a

_

b 20 Ω

+ _

102 V +_

RL

+ Vab

RL

a 16 V

a

b

Step 3 (2 source transformations)

60 Ω

0.4 A Figure 2-31: Circuit evolution for Example 2-14.

section introduces a new circuit-simplification tool—known as the Wye–Delta (Y–) transformation—for dealing specifically with such a circuit arrangement.

R0

1 R2

R1 V0

+ _

R3 3

4 R4

R5 2

Figure 2-32: No two resistors of this circuit share the same current (connected in series) or voltage (connected in parallel).

To that end, let us start by considering the Y and  circuit segments shown in Fig. 2-33(a) and (b), respectively. Let us assume that the same external circuit is connected to the Y and  circuits at nodes 1, 2, and 3. Our task is to develop a set of transformation relations between the resistor set (R1 , R2 , R3 ) of the Y circuit and the resistor set (Ra , Rb , Rc ) of the  circuit that will allow us to replace the Y circuit with the  circuit (or vice versa) without affecting the terminal characteristics (currents and voltages) at nodes 1, 2, and 3. That is, from the standpoint of the external circuit, the Y and  circuits should behave equivalently. The standard procedure employed in deriving the transformation relations is to (a) set one node as an open circuit (i.e.,

“book” — 2015/5/4 — 6:57 — page 82 — #33

82

CHAPTER 2

1

2 R1

c

2-4.1

Rc

1

 →Y Transformation

2 Solution of the preceding set of equations provides the following expressions for R1 , R2 , and R3 :

R2 Rb

R3 3 Y circuit (a)

Ra

3 Δ circuit (b)

R1 =

Rb Rc Ra + R b + R c

(2.42a)

R2 =

R a Rc Ra + R b + R c

(2.42b)

R3 =

R a Rb Ra + R b + R c

(2.42c)

Figure 2-33: Y– equivalent circuits.

not connected to an external circuit), (b) derive an expression for the resistance between the other two nodes (as if a voltage source were connected between them) of theY circuit, (c) follow the same procedure for the  circuit, and then (d) equate the expressions obtained in steps (b) and (c). For example, with node 3 open-circuited, theY circuit reduces to just two in-series resistors R1 and R2 , in which case the resistance between nodes 1 and 2 is simply R12 = R1 + R2

(Y-circuit).

(2.39)

Repeating the procedure for the  circuit (again with node 3 not connected to the external circuit) leads to a configuration between nodes 1 and 2 consisting of Rc in parallel with the series combination of Ra and Rb . Hence, R12 =

Rc (Ra + Rb ) Ra + R b + R c

(-circuit).

(2.40)

Upon equating the expressions for R12 given by Eqs. (2.39) and (2.40), we have R1 + R2 =

RESISTIVE CIRCUITS

Rc (Ra + Rb ) . Ra + R b + R c

(2.41a)

Note the symmetry associated with the form of these expressions:  R1 of the Y circuit, which is connected to node 1, is given by an expression (Eq. (2.42a)) whose numerator is the product of the two resistors connected to node 1 in the  circuit, namely Rb and Rc . The same form of symmetry applies to R2 and R3 .  The transformation represented by the three parts of Eq. (2.42) enables us to replace the  circuit with a Y circuit without having any impact on the external circuit.

2-4.2 Y→  Transformation When applied in the reverse direction, from Y to , the associated transformation relations are given by the following expressions.

Ra =

R1 R2 + R2 R3 + R1 R3 R1

(2.43a)

Rb =

R1 R2 + R2 R3 + R1 R3 R2

(2.43b)

Rc =

R1 R2 + R2 R3 + R1 R3 R3

(2.43c)

When applied to the other two combinations of nodes, the foregoing procedure leads to: Ra (Rb + Rc ) R2 + R3 = Ra + R b + R c

(2.41b)

and Rb (Ra + Rc ) R 1 + R3 = . Ra + R b + R c

(2.41c)

“book” — 2015/5/4 — 6:57 — page 83 — #34

2-4 WYE–DELTA (Y–) TRANSFORMATION

83

For this transformation, the symmetry is as follows:  Ra of the  circuit, which is connected between nodes 2 and 3, is given by an expression (Eq. (2.43a)) whose denominator is R1 , the resistor connected to node 1 of the Y circuit. This form of symmetry also applies to Rb and Rc .  When we started our examination of theY– transformation, we referred to Fig. 2-32. Returning to that figure, we note that the circuit contains two obvious  circuits, namely R1 –R2 –R3 and R3 –R5 –R4 , as well as two not-so-obvious Y circuits: R1 –R3 –R4 and R2 –R3 –R5 . To demonstrate that those two combinations are indeed Y circuits, we have redrawn the circuit in the form shown in Fig. 2-34(a) where we stretched nodes 1 and 2 from single points into two horizontal lines. Electrically, we did not change the circuit whatsoever.

Figure 2-34(b) depicts another rendition of the same circuit. In this case, the Y circuit given by R1 –R3 –R4 resembles a sideways T rather than aY, and the  circuit given by R1 –R3 –R2 resembles a . Hence, it is not surprising that the Y– transformation is oftentimes called the T– transformation. It is instructive to note that the shape in which a circuit is drawn is irrelevant electrically; what does matter is how the branches are connected to the nodes.

2-4.3

Balanced Circuits

If the resistors of the  circuit are all equal, the circuit is said to be balanced (because the three resistors will have equal voltages across them and equal currents through them), as a result of which the Y circuit will also be balanced and will have equal resistors given by

R0

1

Ra 3

(if Ra = Rb = Rc ),

(2.44a)

Ra = Rb = Rc = 3R1 (if R1 = R2 = R3 ).

(2.44b)

R1 = R2 = R3 =

1 and conversely

R2

R1 V0

R3

+

+ -_

3

4

R4

R5

2

Example 2-15: Applying Y– Transformation

2 (a)

R0

1

1

R1 V0

R2 R3

+

+ -_

3

Simplify the circuit in Fig. 2-35(a) by applying the Y– transformation so as to determine the current I . Solution: Noting the symmetry rules associated with the transformation, the  circuit connected to nodes 1, 3, and 4 can be replaced with a Y circuit, as shown in Fig. 2-35(b), with resistances 24 × 36 = 12 , 24 + 36 + 12 24 × 12 R2 = = 4 , 24 + 36 + 12 R1 =

4

R4

R5 and

2

2

(b) Figure 2-34: Redrawing the circuit of Fig. 2-32 to resemble (a) Y and (b) T and subcircuits.

R3 =

36 × 12 = 6 . 24 + 36 + 12

Next, we add the 4  and 20  resistors in series, obtaining 24  for the right branch of the trapezoid. Similarly, the left branch combines into 12  and the two in-parallel branches reduce to a resistance equal to (24 × 12)/(24 + 12) = 8 . When added

“book” — 2015/5/4 — 6:57 — page 84 — #35

84

CHAPTER 2

I



Concept Question 2-10: When is theY– transformation

+

used? Describe the inherent symmetry between the resistance values of the Y circuit and those of the  circuit. (See )

1

36 Ω 100 V

RESISTIVE CIRCUITS

24 Ω 12 Ω

+ _-

3

4 6Ω

Original circuit

(a)

Concept Question 2-11: How are the elements of a

balanced Y circuit related to those of its equivalent  circuit? (See )

20 Ω 2

Exercise 2-9: For each of the circuits shown in Fig. E2.9,

determine the equivalent resistance between terminals (a, b).

I



1

a Req

R1 = 12 Ω

+

R3 = 6 Ω

10 Ω

(a)

R2 = 4 Ω

3

4 6Ω

10 Ω

b

c 100 V + _

10 Ω

20 Ω

a

10 Ω

Req

10 Ω 10 Ω

b (b)

After ∆

Y transformation 2

(b) Figure E2.9 Answer: (a) Req = 15 , (b) Req = 0. (See

I

)

+

100 V

(c)

+ _

25 Ω

Final circuit Figure 2-35: Example 2-15 circuit evolution.

to the 5  and 12  in-series resistances, this leads to the final circuit in Fig. 2-35(c). Hence, I=

100 = 4 A. 25

2-5 The Wheatstone Bridge Developed initially by Samuel Christie (1784–1865) in 1833 as an accurate ohmmeter for measuring resistance, the Wheatstone bridge subsequently was popularized by Sir Charles Wheatstone (1802–1875), who used it in a variety of practical applications. Today, the Wheatstone-bridge circuit is integral to numerous sensing devices, including strain gauges, force and torque sensors, and inertial gyros. The reader is referred to Technology Brief 3 for an illustrative example. The Wheatstone-bridge circuit shown in Fig. 2-36 consists of four resistors: two fixed resistors (R1 and R2 ) of known values,

“book” — 2015/5/4 — 6:57 — page 85 — #36

2-5 THE WHEATSTONE BRIDGE

85

V0

V0

R1

V0

+ −

V1

R2 Ia

Ra R3

R V0

V2

Ammeter

+ −

R

V1

R

Rx

Vout

Figure 2-36: Wheatstone-bridge circuit containing an adjustable variable resistor R3 and an unknown resistor Rx . When R3 is adjusted to make Ia = 0, Rx is determined from Rx = (R2 /R1 )R3 .

V2

Vout

V0 ≈ 4



R R



R + ΔR Flexible resistor

Figure 2-37: Circuit for Wheatstone-bridge sensor.

from which we have an adjustable resistor R3 whose value also is known, and a resistor Rx of unknown resistance. A dc voltage source V0 is connected between the top node and ground, and an ammeter is connected between nodes 1 and 2. The standard procedure for determining Rx starts by adjusting R3 so as to make Ia = 0.

 Rx =

R2 R1

 R3

(balanced condition).

(2.47)

Example 2-16: Wheatstone-Bridge Sensor

 The absence of current flow between nodes 1 and 2, called the balanced condition, implies that V1 = V2 .  From voltage division, V1 = R3 V0 /(R1 + R3 ), V2 = Rx V0 /(R2 + Rx ). Hence, Rx V0 R 3 V0 = . R1 + R 3 R2 + R x

and

(2.45)

A balanced bridge also implies that the voltages across R1 and R2 are equal, R 2 V0 R1 V0 = . R1 + R 3 R2 + R x Dividing Eq. (2.45) by Eq. (2.46) leads to R3 Rx = , R1 R2

(2.46)

A special version of the Wheatstone bridge (Fig. 2-37) is configured specifically for measuring small deviations from a reference condition. An example of a reference condition might be a highway bridge with no load on it. A strain gauge employing a high-sensitivity flexible resistor can measure the small deflection in the bridge surface caused by the weight (force) of a car or truck when present on it. As the force deflects the surface of the bridge to which the resistor is attached, the resistor stretches in length, causing its resistance to increase from a nominal value R (under no stress) to R + R. The other three resistors in the Wheatstone-bridge circuit are all identical and equal to R. Thus, when no vehicles are present on the bridge, the circuit is in the balanced condition. Develop an approximate expression for Vout (the output voltage between nodes 1 and 2) for R/R 1. Solution: Voltage division gives V1 =

V0 V0 R = R+R 2

“book” — 2015/5/4 — 6:57 — page 86 — #37

86

CHAPTER 2

and

RESISTIVE CIRCUITS

I V0 (R + R) V0 (R + R) = . V2 = R + (R + R) 2R + R

I0

Current source I0 Resistor R

Hence, V0 (R + R) V0 − 2R + R 2 2V0 (R + R) − V0 (2R + R) = 2(2R + R) V0 R V0 R = = . 4R + 2 R 4R(1 + R/2R)

Vout = V2 − V1 =

Since R/R 1, ignoring the second term in the denominator would incur negligible error. Such an approximation leads to

Vout

V0 ≈ 4



R R

,

(2.48)

Concept Question 2-12: What is a Wheatstone bridge

)

Concept Question 2-13: What is the balanced condition

in a Wheatstone bridge? (See

1 R

Voltage source V0 V0

V

Figure 2-38: I –V relationships for a resistor R, an ideal voltage source V0 , and an ideal current source I0 .



providing a simple linear relationship between the change in resistance R and the output voltage Vout .

used for? (See

slope =

)

Many very useful circuit elements do not have linear i–υ relationships. Consider Fig. 2-39(a). A realistic voltage source is connected to a load RL at terminals (a, b). Note that the resistance value of the source resistor Rs is much smaller than that of the load (1  versus 1 k). It is typical of a well-designed voltage source to have a small source resistor so as to minimize the voltage drop across it. The switch simulates an accidental short circuit. Application of KVL to the loop in Fig. 2-39(a) (with the switch in the open position) leads to Is =

Exercise 2-10: If in the sensor circuit of Fig. 2-37, V0 = 4 V and the smallest value of Vout that can be measured reliably is 1 μV, what is the corresponding accuracy with which (R/R) can be measured? Answer: 10−6 or 1 part in a million. (See

2-6.1 The Fuse: A Simple Nonlinear Element

)

2-6 Application Note: Linear versus Nonlinear i–υ Relationships Ideal resistors and voltage and current sources are all considered linear elements; the relationship between the current and the voltage across any one of them is described by a straight line. The i–υ relationships plotted in Fig. 2-38 for the current source, the voltage source, and the resistor have slopes of 0, ∞, and 1/R, respectively.

Vs 100 = ≈ 0.1 A Rs + R L 1 + 1000

(switch open).

If, accidentally, a short circuit were to be introduced across terminal (a, b), which is represented schematically by the closing of the SPST switch, the current Is will flow entirely through the short circuit, resulting in Is =

Vs = 100 A! Rs

(switch closed).

This is a very large current. Many household wires would begin to overheat and melt off their insulation at such high currents. It is precisely for this reason that the fuse (and later, the breaker) came into heavy use in power-distribution circuits [Fig. 2-39(b)]. The i–υ curve for a fuse, shown in [Fig. 2-39(c)], is decidedly nonlinear: Above a certain current level, the fuse will cease to allow more current to pass through it, acting like a current limiter. The physical device contains a small metal wire that is designed to melt away at a specific current level

“book” — 2015/5/4 — 6:57 — page 87 — #38

2-6 APPLICATION NOTE: LINEAR VERSUS NONLINEAR I –υ RELATIONSHIPS

Is + _

Rs = 1 Ω Vs = 100 V

87

ID

a Accidental short circuit

RL = 1 kΩ

Anode (p-type)

VF

VD

RD

Cathode (n-type) Load b (a) Accidental short circuit represented by a switch

Source

(a) Diode symbol

(b) Realistic diode model

ID Is + _

Rs

If

Vs = 100 V

Fuse Vf

a RL

Accidental short circuit

Knee voltage = 0 VD

b Load (b) Fuse to protect voltage source

Source with fuse

(c) i−υ of an ideal diode ID

If

slope = RD Real diode response

Overcurrent limit

Vf

(d) i−υ of a real diode

(c) i−υ characteristic for a fuse prior to opening Figure 2-39: Use of a fuse to protect a voltage source.

(called its overcurrent), thereby becoming an open circuit and preventing large currents from flowing through the circuit. Note that Fig. 2-39(b) does not explain the fuse’s time-dependent behavior; it describes the fuse’s behavior only until the moment at which the current exceeds the overcurrent. After that, the fuse just looks like an open circuit. Fuses also are rated for several other important characteristics such as how fast they can respond. Ultra-fast fuses can trip in milli- to micro-seconds. Another important attribute is the maximum voltage it can sustain across its terminals. Note that in Fig. 2-39(b), once the fuse assumes the role of an open circuit, the voltage across it becomes Vs . If this voltage is too high, arcing and sparks might develop between the terminals (we know from physics that a large-enough voltage in air will break down the air molecules, causing them to conduct and generate a bright spark). Clearly, that is an important rating factor to keep in mind when selecting a fuse.

VD

VF

ID Approximate practical diode response VD Forward voltage VF (e) Approximate diode response Figure 2-40: pn-junction diode schematic symbol and i–υ characteristics.

2-6.2 The Diode: A Solid-State Nonlinear Element The diode is a mainstay of solid-state circuits. Its circuit schematic symbol is shown in Fig. 2-40(a) with VD as the voltage across the diode, defined such that the (+) side is at the anode terminal of the diode and the (−) side at its

“book” — 2015/5/4 — 6:57 — page 88 — #39

88 cathode terminal. There are many types of diodes, including the basic pn-junction diode, the Zener and Schottky diodes, and the ubiquitous light-emitting diode (LED) used in consumer electronics. A brief introduction of the LED was made earlier in Section 2-1.4, and a more detailed overview of its operation is provided in Technology Brief 5. For the present, we will limit our discussion to the pn-junction diode, commonly referred to simply as the diode. The pn diode consists of a p-type semiconductor placed in contact with an n-type semiconductor, thereby forming a junction. The p-type material is so named because the impurities that have been added to its bulk material result in a crystalline structure in which the available charged carriers are predominantly positive charges. The opposite is true for the n-type material; different types of impurities are added to the bulk material, as a result of which the predominant carriers are negative charges (electrons). In the absence of a voltage across the diode, the two sets of carriers diffuse away from each other at the edge of the junction, generating an associated builtin potential barrier (voltage), called the forward-bias voltage or offset voltage VF . The main use of the diode is as a one-way valve for current. Figure 2-40(c) displays the i–υ relationship for an ideal diode, which conveys the following behavior:  Current can flow through the diode from the (+) terminal to the (−) terminal unimpeded, regardless of its magnitude, but it cannot flow in the opposite direction.  In other words, an ideal diode looks like a short circuit for positive values of VD and like an open circuit for negative values of VD . These two states are called forward bias and reverse bias, respectively. When a positive-bias voltage exceeding VF is applied to the diode, the potential barrier is counteracted, allowing the flow of current from p to n (which includes positive charges flowing in that direction as well as negative charges flowing in the opposite direction). On the other hand, if a negative-bias voltage is applied to the diode, it adds to the potential barrier, further restricting the flow of charges across the barrier and resulting in no current flow from n to p. The voltage level at which the diode switches from reverse bias to forward bias is called the knee voltage or forward-bias voltage. For the ideal diode, VF = 0 and the knee is at VD = 0, which means that the forward-bias segment of its i–υ characteristic is aligned perfectly along the ID axis, as shown in Fig. 2-40(c). Real diodes differ from the ideal diode model in two important respects: (1) the knee in the curve is not at VD = 0, and (2) the diode does not behave exactly like a perfect short circuit when in forward bias nor like a perfect open circuit

CHAPTER 2

RESISTIVE CIRCUITS

when in reverse bias. Figure 2-40(d) shows a realistic diode i–υ curve. Note how nonlinear a real diode really is! For many electrical engineering applications, however, the nonlinearities are not so important, and the approximate ideal-like diode model shown in Fig. 2-40(e) is quite sufficient. The only difference between the ideal diode model of Fig. 2-40(c) and the approximate diode model of Fig. 2-40(e) is that in the latter the transition from reverse to forward bias occurs at a non-zero, positive value of VD , namely the forward-bias voltage VF . For a silicon pn-junction diode, a typical value of VF is 0.7 V and a realistic model is shown in Fig. 2-40(b). A typical value of RD is 10–20 . We always should remember that VF is a property of the diode itself, not of the circuit it is a part of. Example 2-17: Diode Circuit

The circuit in Fig. 2-41 contains a diode with VF = 0.7 V. Determine ID , assuming RD to be negligibly small. Solution: Initially, we do not know whether the diode is forward biased or reverse biased. We will first assume it is forward biased in order to compute ID . Then, if it turns out that ID is positive, our assumption will have been validated, but if ID is negative, we will conclude that the diode is reverse biased and no current flows through the circuit. Application of KVL around the loop gives −Vs + ID R + VD = 0. If the diode is forward biased, VD = 0.7 V, which leads to ID =

Vs − VD 5 − 0.7 = = 43 mA. R 100

The positive sign of ID confirms our assumption that the diode is indeed forward biased. As an interesting aside, one could use this circuit to control the current through a light-emitting diode (LED). As explained in Technology Brief 5, the amount of light emitted by an LED (i.e., how bright it appears) is proportional directly to the current ID passing through it when it is forward biased. By using the circuit

VR Vs = 5 V

+ _

ID

R = 100 Ω

Figure 2-41: Diode circuit of Example 2-17.

VD

“book” — 2015/5/4 — 6:57 — page 89 — #40

2-6 APPLICATION NOTE: LINEAR VERSUS NONLINEAR I –υ RELATIONSHIPS in Fig. 2-41 and choosing an appropriate value for R, we can build a circuit that forward biases an LED and controls its brightness. Example 2-18: Square-Wave Waveform

The circuit in Fig. 2-42 contains two diodes, both with VF = 0.7 V. The waveform of the voltage source consists of a single cycle of a square wave. Generate plots for i1 (t) and i2 (t). Ignore RD for both diodes. Solution: Again, we will use the diode model of Fig. 2-40(b). From the analysis of Example 2-17, we

υa + _

υs(t)

i1

+ _

D1

R1

53 Ω

_ + R2

concluded that if the voltage across a series combination of a diode and a resistor exceeds VF of the diode (with the + polarity of the voltage coinciding with the + side of the diode), current will flow through the series combination, but if the voltage is negative, no current will flow through the diode. For the first half of the source voltage cycle, υa , the voltage across the series combination (D1 , R1 ) is positive at 6 V. Hence, i1 (t) =

υa − 0.7 6 − 0.7 = = 0.1 A R1 53

106 Ω

i2 (t) = 0

for 0 ≤ t ≤ 1 s.

The opposite behavior occurs during the second half of the cycle of υs (t), diode D2 will conduct current through it, but diode D1 will not. Hence, i1 (t) = 0 i2 (t) =

(a) Diode circuit

for 0 ≤ t ≤ 1 s.

But for series combination (D2 , R2 ), no current will flow through diode D2 because the polarity of υa is opposite of that of the diode. Hence,

i2 D2

89

for 1 ≤ t ≤ 2 s,

6 − 0.7 6 − 0.7 = 0.05A = R2 106

for 1 ≤ t ≤ 2 s.

The combined results are displayed in Fig. 2-42(c).

υs (t)

Concept Question 2-14: What is the overcurrent of a

6V

fuse? (See

t (s)

2

1

)

Concept Question 2-15: Why does a pn-junction diode

have a non-zero forward-bias voltage VF? (See

)

−6 V Exercise 2-11: Determine I in the two circuits of Fig. E2.11. Assume VF = 0.7 V for all diodes.

(b) Source voltage waveform

2 kΩ 0.1 A

i1(t)

0.05 A 1

3 kΩ

12 V

i2(t) 2

I

I

2 kΩ

3 kΩ

12 V

t (s)

(c) Current waveforms Figure 2-42: Diode circuit and waveforms of Example 2-18.

(a)

(b) Figure E2.11

Answer: (a) I = 2.12 mA, (b) I = 0. (See

)

“book” — 2015/5/4 — 6:57 — page 90 — #41

90

TECHNOLOGY BRIEF 5: LIGHT-EMITTING DIODES (LEDS)

Technology Brief 5 Light-Emitting Diodes (LEDs)

Longer leg

How LEDs Are Made

Anode

LEDs are a specific type of the much larger family of semiconductor diodes, whose basic behavior we discussed earlier in Section 2-6. When a voltage is applied in the forward-biased direction across an LED, current flows and photons are emitted (Fig. TF5-1). This occurs because as electrons surge through the diode material, they recombine with charge carriers in the material and release energy in the form of photons (quanta of light). The energy of the emitted photon (and hence the wavelength/color) depends on the type of material used to make the diode. For example, a diode made of indium gallium aluminum phosphide (InGaAlP) emits red light, while a diode made from gallium nitride (GaN) emits bluish light. Extensive research over many decades has yielded materials that can emit photons at practically any wavelength from the infrared through ultraviolet (Fig. TF5-2). Various “tricks” have also been employed to modify the emitted light after emission. To make white light diodes, for example, certain blue light diodes can be coated with crystal powders which convert the blue light into a broad-spectrum “white” light. Other coatings such as quantum dots are still the subject of today’s research. In a traditional package, the LED transmits light in a hemispherical pattern, but numerous other light-focused packaging methods are available that can focus the light in virtually any way imaginable. LEDs can be focused using highly reflective coatings to intensify their light for higher power applications.

Lens Light

Metal leads Cathode Light-emitting semiconductor diode Figure TF5-1: Basic configuration of an LED. In addition to semiconductor LEDs, a newer class of devices called organic light emitting diodes (OLEDs) are the subject of intense research efforts. OLEDs operate in a manner that is analogous to conventional LEDs, but are made from organic molecules (often polymers). Because OLEDs are lighter weight than conventional LEDs and can be made to be flexible, they have the potential to revolutionize handheld and lightweight displays, such as those used in phones, PDAs and flexible screens. Imagine a flexible contact lens that could allow you to see a heads-up display or augmented reality!

LED Advantages LEDs have several major attributes that have made them a key element of many applications. First, they can be

de phi e hos allium sphid p ic m g e o n d s h i e o r p s t u ni ph ar Indi num lium lium lium mi Gal alu Gal G al

um

min

de phi

alu lium e l a G nid e ars

Intensity

Material composition

Epoxy case

350

400

UV

450

500

550

600

650

Wavelength (nm)

700

750

800

850

900

950

Infrared

Figure TF5-2: Emission spectra of LEDs made of different material composites.

1000

“book” — 2015/5/4 — 6:57 — page 91 — #42

TECHNOLOGY BRIEF 5: LIGHT-EMITTING DIODES (LEDS)

91

FigureTF5-4: LED eyelashes can be worn in many colors, and can be made to turn on or off with a tip of the head. (Credit: Soomi Park.)

Figure TF5-3: LED-lit building.

produced in a wide variety of wavelengths from infrared through ultraviolet. Targeted or broad spectra can also be produced, making them applicable to virtually any optical application. Second, they are energy efficient. An incandescent lightbulb uses 80% of its energy for heat and 20% to produce light. LEDs use only about 20% of their energy for heat and 80% for light. This also makes them cool, requiring less energy to remove the excess heat. Third, they are manufactured in a huge array of colors, sizes, shapes, designs, and more. They are affordable (not yet less expensive than incandescent bulbs in the initial purchase price, but definitely less expensive over the lifetime of the bulb). Fourth, they last longer (often > 100k hours) than incandescent bulbs, which is particularly important in hard-to-reach applications. Fifth, they can be integrated directly into semiconductor circuits, printed circuit boards, and light-focusing packages. Various combinations of these advantages are key to the following broad range of applications of LEDs.

LEDs for Lighting In an era where energy efficiency matters financially, environmentally, and practically, LEDs have become a popular mainstay in home and office lighting, street lighting and consumer products from home appliances and toys to high-efficiency tail lights for cars and flashlights. Of growing importance is the replacement of traditional incandescent bulbs with LEDs in homes and buildings (Fig. TF5-3), because of their energy efficiency.

But lighting is more than just enabling us to see at night. LEDs can be used in horticulture to efficiently target ideal wavelengths for plant growth, and exposing produce to certain wavelengths of light can help it ripen on demand, or can extend its ripened shelf life. UV LEDS are being explored to enhance development of polyphenol, which are believed to have antioxidant qualities, in growth of green, leafy vegetables. LEDs provide high visibility bike lights, safety vests, tennis shoes, and more. They are also used artistically for decoration and advertising on buildings and signs, woven into clothes often augmented by plastic fiber optic threads (e.g., Philips Research Lumalive textiles), or even worn with LED eyelashes (see Fig. TF5-4)!

LEDs for Medical Applications LEDs are used for a variety of medical applications. One particularly important application is the pulse oximeter (Fig. TF5-5), which measures blood oxygen level and pulse rate. Oxygenated blood absorbs light at 660 nm (red light), whereas deoxygenated blood absorbs light at 940 nm (infrared). Pulse oximeters use two LEDs, one at 600 nm and another at 940 nm, which are arranged to transmit through a translucent section of the body such as the finger or ear lobe. Two associated light collecting sensors are placed on the opposite side to measure the amount of each wavelength that is transmitted through the body. The ratio of the red and infrared light indicates how much oxygen is in the blood. To insure that the received light signals are actually from the blood, the measurement is made over several seconds (several pulses), focusing in on the pulsing blood rather than the static surrounding tissues.

“book” — 2015/5/4 — 6:57 — page 92 — #43

92

TECHNOLOGY BRIEF 5: LIGHT-EMITTING DIODES (LEDS)

Figure TF5-5: Pulse oximeter used to measure blood oxygen content.

LEDs are also used to treat many superficial (skin) conditions. Red light in the range of 600–950 nm can be used to treat acne, rosacea, and wrinkles. The red light works by stimulating the mitochondria in the skin to make older cells behave like younger cells. Blue-light therapy in the 405–420 nm range is used for acne treatments and “anti-aging” skin therapies because of its ability to stimulate collagen in the skin. Green to yellow light (532– 595 nm) can reduce skin redness (rosacea). Combining LED light sources with topical drug treatments that are photoactivated may be used to treat a variety of skin conditions including skin cancer and pre-cancer. LEDs are also used extensively in dentistry. Blue LEDs can be used to cure (harden) polymer composite materials used for fillings. The rate at which the filling material cures is proportional to the power carried by the LED light, so high power LEDs are used to speed up the curing process.

Ultraviolet (UV) LEDs The UV range provides a wealth of applications, and low-cost high-power UV LEDs are enabling many of these applications. Inks (printing), adhesives and coatings are often cured with LEDs in the UV range (primarily 395 nm, 385 nm or 365 nm). UV LED flashlights are used to detect fraudulent identification (at the airport, for example) and currency. UV-LEDs are used extensively in forensic analysis and drug discovery. In the lower UV spectral range (100–280 nm) LEDs sterilize air and water by breaking up the DNA and RNA of

Figure TF5-6: Large LED display.

microorganisms and preventing their reproduction. For example, 275 nm is believed to be the most effective wavelength for eradicating pathogens such as E-coli in water. LEDs in this range are also used for spectroscopic and fluorescence measurements and for chemical and biological detectors.

LED Displays LEDs, with their wide range of colors, efficiency, low cost, flexibility, low profile and light weight, are ideal for both handheld displays and much larger displays (such as billboards and signage, as shown in Fig. TF5-6). Some LED displays use edge lighting where LEDs shine light across the screen (allowing the display to be thinner than traditional screens but not improving picture quality). Others use RGB LEDs.These LEDs use a common anode but have separate cathodes for red, green and blue LEDs (making the composite a 4-pin LED). They can be made to generate light with almost any color, depending on the voltages applied across the combination of RGB pins. This greatly enhances picture color. RGB LEDs can also be dimmed independently and instantly (giving a more dynamic picture, especially great “black” levels for dark scenes). The flexibility and bendability of OLEDs promise new, creative options for the next generation of TVs and smart phones—can you imagine rolling your TV up like a poster and carrying it with you anywhere? Or wearing it? Or . . . ?

“book” — 2015/5/4 — 6:57 — page 93 — #44

2-6 APPLICATION NOTE: LINEAR VERSUS NONLINEAR I –υ RELATIONSHIPS

Mechanical load P (N/m2)

No load

Δx R

R + ΔR

Figure 2-43: The resistance of a piezoresistor changes when mechanical stress is applied.

2-6.3

Piezoresistor Circuit

According to Technology Brief 4, if we apply a force on a resistor along its axis (Fig. 2-43), the resistance changes from R0 , which is the resistance with no stress (pressure) applied, to R as R = R0 + R,

of such a sensor. As a reference, a finger can apply about 50 N of force across an area of 1 cm2 (10−4 m2 ), which is equivalent to a pressure P = 5 × 105 N/m2 . If the piezoresistor is made of silicon with α = −1 × 10−9 m2 /N and if the dc source in the Wheatstone bridge is V0 = 1 V, Eq. (2.51) yields the result that Vout = −125 μV, which is not impossible to measure but quite small nevertheless. How then are such pressure sensors used? The answer is simple: We need a mechanism to amplify the signal. We can do so electronically by feeding Vout into a highgain amplifier, or we can amplify the mechanical pressure itself before applying it to the piezoresistor. The latter approach can be realized by constructing the piezoresistor into a cantilever structure, as shown in Fig. 2-44 (a cantilever is a fancy name for a “diving board” with one end fixed and the other free). Deflection of the cantilever tip induces stress at the base of the cantilever near the attachment point. If properly designed, the cantilever—which usually is made of silicon or metal—can amplify the applied stress by several orders of magnitude, as we see in the following example.

(2.49)

and the deviation R is given by R = R0 αP,

93

Example 2-19: A Realistic Piezoresistor Sensor

(2.50)

where α is a property of the material that the resistor is made of and is called its piezoresistive coefficient, and P is the mechanical stress applied to the resistor. The unit for P is newtons/m2 (N/m2 ) and the unit for α is the inverse of that. Compression decreases the length of the resistor and increases its cross section, so in view of Eq. (2.2), which states that the resistance of a longitudinal resistor is given by R = ρ/A, the consequence of a compressive force—namely reduction in  and increase in A—leads to a reduction in the magnitude of R.

When a force F is applied on the tip of a cantilever of width W , thickness H , and length L (as shown in Fig. 2-44) the corresponding stress exerted on the piezoresistor attached to the cantilever base is given by P=

FL . WH2

(2.52)

Determine the output voltage of a Wheatstone-bridge circuit if F = 50 N, V0 = 1 V, the piezoresistor is made of silicon, and the cantilever dimensions are W = 0.5 cm, H = 0.5 mm, and L = 1 cm. Solution: Combining Eqs. (2.51) and (2.52) gives

 Hence, for compression, R is negative, requiring that α in Eq. (2.50) be defined as a negative quantity.  If a piezoresistor is integrated into a Wheatstone-bridge circuit (as in Fig. 2-37), such that all three other resistors are given by R0 , the expression for the voltage output given by Eq. (2.48) becomes   V0 R V0 Vout = = αP. (2.51) 4 R0 4 Since V0 and α are both constants, the linear relationship between the applied stress P and the output voltage Vout makes the piezoresistor a natural sensor for detecting or measuring mechanical stress. However, we should examine the sensitivity

V0 FL α· 4 WH2 1 50 × 10−2 = × (−1 × 10−9 ) × −3 4 (5 × 10 ) × (5 × 10−4 )2

Vout =

= −0.1 V. The integrated piezoresistor–cantilever arrangement generates an output voltage whose magnitude is on the order of 800 times greater than that generated by pressing on the resistor directly! Concept Question 2-16: Does compression along the current direction increase or decrease the resistance? Why? (See )

“book” — 2015/5/4 — 6:57 — page 94 — #45

94

CHAPTER 2

RESISTIVE CIRCUITS

Piezoresistor Rest position P

Force F

W

H L Deflected position

Figure 2-44: A cantilever structure with integrated piezoresistor at the base. Concept Question 2-17: Why are piezoresistors placed

In this section, you will learn how to:

at the base of cantilevers and other deflecting structures? (See )

• Set up and analyze a simple dc circuit in Multisim.

Exercise 2-12: What would the output voltage associated

with the circuit of Example 2-18 change to, if the cantilever thickness is reduced by a factor of 2? Answer: Vout = −0.4 V. (See

)

2-7 Introducing Multisim Multisim 13 is the latest edition of National Instrument’s SPICE simulator software. SPICE, originally short for Simulation Program with Integrated Circuit Emphasis, was developed by Larry Nagel at the University of California, Berkeley, in the early 1970s. It since has inspired and been used in many academic and commercial software packages to simulate analog, digital, and mixed-signal circuits. Modern SPICE simulators like Multisim are indispensable in integrated circuit design; ICs are so complex that they cannot be built and tested on a breadboard ahead of production (see Technology Brief 7). With SPICE, you can draw a circuit from a library of components, specify how the components are connected, and ask the program to solve for all voltages and currents at any point in time. Modern SPICE packages like Multisim include very intuitive graphic user interface (GUI) tools that make both circuit design and analysis very easy. Multisim allows the user to simulate a laboratory experience on his/her computer ahead of actually working with real components.

• Use the Measurement Probe tool to quickly solve for voltages and currents. • Use the Analysis tools for more comprehensive solutions. We will return to these concepts and learn to apply many other analysis tools throughout the book. Appendix C provides an introduction to the Multisim Tutorial available on the book website http://c3.eecs.umich.edu/. The Tutorial is a useful reference if you have never used Multisim before. When defining menu selections starting from the main window, the format Menu → Sub-Menu1 → Sub-Menu2 will be used.

2-7.1

Drawing the Circuit

After installing and running Multisim, you will be presented with the basic user interface window, also referred to as the circuit window or the schematic capture window (see Multisim Tutorial on the book website). Here, we will draw our circuits much as if we were drawing them on paper. Placing resistors in the circuit Components in Multisim are organized into a hierarchy going in a descending general order from Database → Group → Family → Component. Every component that you use in Multisim will fit into this hierarchy somewhere. Place → Component opens the Select a Component window. (Ctrl-W is the shortcut key for the placecomponent command. Multisim has many shortcut keys,

“book” — 2015/5/4 — 6:57 — page 95 — #46

2-7

INTRODUCING MULTISIM

95

Figure 2-45: Multisim screen for selecting and placing a resistor. and it will be worthwhile for you to learn some of the basic ones to improve your efficiency in creating and testing circuits.) Choose Database: Master Database and Group: Basic in the pulldown menus.

Also, by double-clicking on a specific component, you can access many details of the component model and its values. For now, it is sufficient to know that the Resistance value can be altered at any time through the Value menu. Placing an independent voltage source

Now select Family: RESISTOR. You should see a long list of resistor values under Component and the schematic symbol for a resistor (Fig. 2-45). Note that the Family menu contains other components like inductors, capacitors, potentiometers, and many more. We will use these in later chapters. Scroll down and select a 1k value (the units are in ohms) and then click OK. You should see a resistor in the capture window. Before clicking in the window, Ctrl-R allows you to rotate the resistor in the window. Rotate the resistor such that it is vertical and then click anywhere on the window to place it. Repeat this operation; this time place a vertical 100-ohm resistor directly below the first one (as in Fig. 2-46). How to connect them together will be described shortly. Once you are finished placing components, click Close to return to the schematic capture window. Note that the components have symbolic names (R1 and R2) and values displayed next to them (1k and 100).

Just as you did with the resistors, open up the Select a Component window. Choose Database: Master Database and Group: Sources in the pulldown menus. Select Family: POWER SOURCES. Under Component select DC POWER and click OK. Place the part somewhere to the left of the two resistors (Fig. 2-46). Once placed, close the component window, then doubleclick on component V1. Under the Value tab, change the Voltage to 10 V. Click OK. Wiring components together Place → Wire allows you to use your mouse to wire components together with click-and-drag motions (Ctrl-Q is the shortcut key for the wire command). You

“book” — 2015/5/4 — 6:57 — page 96 — #47

96

CHAPTER 2

RESISTIVE CIRCUITS

Simulation toolbar Probe 1 display 1 Node 1

Component name

Node 2

2

Component value

Probe 1 display 0

Form wire corner by clicking here as you drag wire

Ground

Finish dragging wire to R2 to complete circuit

Figure 2-47: Executing a simulation.

2-7.2 Figure 2-46: Adding a voltage source and completing the circuit.

can also enable the wire tool automatically by moving the cursor very close to a component node; you should see the mouse pointer change into a black circle with a cross-hair. Click on one of the nodes of the dc source with the wire tool activated (you should see the mouse pointer change from a black cross to a black circle with a cross hair when you hover it over a node). Additional clicks anywhere in the schematic window will make corners in the wire. Doubleclicking will terminate the wire. Additionally, when not already dragging a wire, double-clicking on any blank spot of the schematic will generate a wire based at the origin of clicking. Wire the components as shown in Fig. 2-46. Add a GROUND reference point as shown in Fig. 2-47. The Ground can be found in the Component list of POWER SOURCES. We now have a resistive divider.

Solving the Circuit

In Multisim, there are two broad ways in which to solve a circuit. The first, called Interactive Simulation, allows you to utilize virtual instruments (such as ohmmeters, oscilloscopes, and function generators) to measure aspects of a circuit in a time-based environment. It is best to think of the Interactive Simulation as a simulated “in-lab” experience. Just as in real life, time proceeds in the Interactive Simulation as you analyze the circuit (although the rate at which time proceeds is heavily dependent on your computer’s processor speed and the resolution of the simulation). The Interactive Simulation is started using the F5 key, the

button, or the

toggle

switch. The simulation is paused using the F6 key, the button, or the

button. The simulation is terminated using

either the button or the toggle switch. The other main way in which to solve a circuit in Multisim is through Analyses. These simulations display their outputs not in instruments, but rather in the Grapher window (which

“book” — 2015/5/4 — 6:57 — page 97 — #48

2-7

INTRODUCING MULTISIM

may produce tables in some instances). These simulations are run for controlled amounts of time or over controlled sweeps of specific variables or other aspects of the circuit. For example, a dc sweep simulates the values of a specified voltage or current in the circuit over a defined range of dc input values. Each of the methods described has its own advantages and disadvantages, and in fact, both varieties can perform many of the same simulations, albeit with different advantages. The choice of method to be used for a given circuit really comes down to your preferences, which will be formed as you gain more experience with Multisim. For the circuit in Fig. 2-47, we wish to solve for the voltages at every node and the currents running through every branch. As you will often see in Multisim, the solution can be obtained using either the Interactive Simulation or through one of the Analyses. We will demonstrate both approaches.

97 must be stopped, not just paused, in order for the DC Operating Point Analysis mode to work.] Under the Output tab, select the two node voltages and the branch current in the Variables in Circuit window. Make sure the Variables in Circuit pull-down menu is set to All Variables. Once selected, click Add and they will appear in the Selected variables for analysis window. Once you have selected all of the variables for which you want solutions, simply click Simulate. Multisim then solves the entire circuit and opens a window showing the values of the selected voltages and currents (Fig. 2-48).

2-7.3

Dependent Sources

Interactive simulation

Multisim provides both defined dependent sources (voltagecontrolled current, current-controlled current, etc.) and a generic dependent source whose definition can be entered as a mathematical equation. We will use this second type in the following example.

Selecting Simulate → Instruments → Measurement Probe allows you to drag and place a measurement probe onto any node in the circuit. (Note that the Instruments menu contains many common types of equipment used in an electronics laboratory.) The Measurement Probe constantly reports both the current running through the branch to which it is assigned and the voltage at that node. Place two probes into the circuit as shown in Fig. 2-47. When placed, by default, the probes should be pointing in the direction shown in Fig. 2-47. If they are not, you can reverse a probe’s direction by right-clicking on it and pressing Reverse Probe Direction. Once the probes are in place, you must run the simulation using the commands for Interactive Simulations. As expected, the current running through both wires is the same since the circuit has only one loop.

Step 1: The dependent sources are established as follows: Place → Component opens the Select a Component window. Choose Database: Master Database and Group: Sources in the pulldown menus. Select Family: CONTROLLED VOLTAGE or CONTROLLED CURRENT. Under Component, select ABM VOLTAGE or ABM CURRENT and click OK. The value of ABM sources (which stands for Analog Behavioral Modeling) can be set directly with mathematical expressions using any variables in the circuit. For information on the variable nomenclature, which may be somewhat confusing, see the Multisim Tutorial on the book website.

I=

10 V1 = = 9.09 mA. R1 + R 2 1000 + 100

The voltage at node 1 is 10 V, as defined by the source. Application of voltage division (Fig. 2-19) gives  V2 =

R2 R1 + R 2



 V1 =

 100 10 = 0.909 V. 1100

DC operating point analysis The circuit also can be solved using Simulate → Analyses → DC Operating Point. This method is more convenient than the Interactive Simulation when solving circuits with many nodes. After opening this window, you can specify which voltages and currents you want solved. [The Interactive Simulation mode

Step 2: Using what you learned in Section 2-7.1, draw the circuit shown in Fig. 2-49 (including the probe at node 2). Step 3: Double-click the ABM CURRENT source. Under the value tab, enter: 3*V(2). The expression V(2) refers to the voltage at node 2. This effectively defines this source as a voltage-controlled current source. Note that when making the circuit, if the node numbering in your circuit differs from that in the example (e.g., if nodes 1 and 2 are switched), then take care to keep track of the differences so that you will use the proper node voltage when writing the equation. To edit or change node labels, double-click any wire to open the Net Window. Under Net name enter the label you like for that node. To write the expression for I1 next to the current source, go to Place → Text, and then type in the expression at a location near

“book” — 2015/5/4 — 6:57 — page 98 — #49

98

CHAPTER 2

RESISTIVE CIRCUITS

Node V1 1

V(1)

Voltage @ node V1 Voltage @ node V2 Current through node V1

V(2) I(v1)

Node V2

2

0

Figure 2-48: Solution window. Referencing currents in arbitrary branches

Figure 2-49: Creating a dependent source.

the current source. [Ctrl-T is the shortcut key for the place-text command.]

Now let us analyze the circuit using the DC Operating Point Analysis. Our goal is to solve for the voltages at every node and the current running through each branch. Remove the probe from the circuit if you still have it in there by clicking on it so it is highlighted and pressing the Delete key. To perform a DC operating point analysis, just as we did earlier in Section 2-7.2, go to Simulate → Analyses → DC Operating Point and transfer all available variables into the Selected variables for analysis window. You should notice that the only variables available are V(1), V(2), and I(v1); if Probe 1 is still connected to your circuit, you should also see I(Probe 1) and V(Probe 1). Where are the other currents, such as the current flowing through R1, the current through R2, or even the current coming out of the dependent source? In Multisim and most SPICE software in general, you can only measure/manipulate currents through a Voltage Source (there are some exceptions, but we will ignore them for now). This is why the current through V1, denoted I(v1), is available but the currents through the other components are not. A simple

“book” — 2015/5/4 — 6:57 — page 99 — #50

2-7

INTRODUCING MULTISIM

trick, however, to obtain these currents is to add a 0 V dc source into the branches where you want to measure current. Do this to your circuit, so that it ends up looking like that shown in Fig. 2-50.

99

Concept Question 2-19: How do you obtain and

visualize the circuit solution? (See

)

Exercise 2-13: The circuit in Fig. E2.13 is called a resistive bridge. How does Vx = (V3 − V2 ) vary with the value of potentiometer R1 ?

1

3

2

0

Figure 2-50: Circuit from Fig. 2-49 adapted to read out the currents through R1, R2, and the dependent source.

Figure E2.13 Answer: (See

You will notice that there are new nodes in the circuit now, but since V2, V3, and V4 are 0 V sources, V(3) = V(4) = V(1) and V(5) = V(2). Go back to the DC Operating Point Analysis window and under the Variables in Circuit window there should now be four currents [I(v1), I(v2), I(v3), and I(v4)] and the five voltages. Highlight all four currents as well as V(1) and V(2) and click Add and then click OK. This will bring up the Grapher window with the results of the analysis. Note that when we analyze the currents through the branches, the current through a voltage source is defined as going into the positive terminal. For example, in source V1, this corresponds to the current flowing from Node 1 into V1 and then out of V1 to Node 0. Concept Question 2-18: In Multisim, how are components placed and wired into circuits? (See )

)

Exercise 2-14: Simulate the circuit shown in Fig. E2.14 and solve it for the voltage across R3 . The magnitude of the dependent current source is V1 /100.

1

2 R2 100 Ω

V1 12 V

10 Ω ABM ABM_CURRENT

0

4 Figure E2.14

Answer: (See

R1

)

3 R3 1 Ω

“book” — 2015/5/4 — 6:57 — page 100 — #51

100

CHAPTER 2

RESISTIVE CIRCUITS

Summary

Concepts • As described by Ohm’s law, the i–υ relationship of a resistor is linear over a specific range (−imax to +imax ); however, R may vary with temperature (thermistors), pressure (piezoresistors), and light intensity (LDR). • Kirchhoff’s current and voltage laws form the foundation of circuit analysis and synthesis. • Two circuits are considered equivalent if they exhibit identical i–υ characteristics relative to an external circuit. • Source transformation allows us to represent a real voltage source by an equivalent real current source, and vice versa.

Mathematical and Physical Models Linear resistor

• A Y circuit configuration can be transformed into a  configuration, and vice versa. • The Wheatstone bridge is a circuit used to measure resistance, as well as to detect small deviations (from a reference condition), as in strain gauges and other types of sensors. • Nonlinear resistive elements include the light bulb, the fuse, the diode, and the light-emitting diode (LED). • Multisim is a software simulation program capable of simulating electric circuits and analyzing their behavior. • A diode is a one-way valve for current. An LED is a diode that also emits light.

Voltage division

R = ρ/A p = i2R N 

Kirchhoff current law (KCL) in = current entering node n

υn = voltage across branch n

R2 + _ υ2 =

in = 0

υs υs

n=1

Current division N 

Kirchhoff voltage law (KVL)

υs

R1 R1 + R2 R2 R1 + R2

R1 + _ υ1 =

+ _

n=1

i1 =

vn = 0

is

R1

R2

i2 =

R2 R1 + R2 R1 R1 + R2

G1 is Geq G2 is is = Geq is =

Resistor combinations N 

In series

Req =

In parallel

 1 1 = Req Ri

Ri

i=1 N

Source transformation Rs

+ υs _

i=1

or Geq =

N  i=1

Gi

is =

υs Rs

Rs

Y– transformation

Table 2-5

Wheatstone bridge (Fig. 2-37)

υout ≈

V0 4



R R



“book” — 2015/5/4 — 6:57 — page 101 — #52

PROBLEMS

Important Terms American Wire Gauge ammeter Analyses balanced balanced condition basic user interface breaker circuit equivalence circuit window conductance conductivity conductor current divider dielectric diode equivalent resistor forward bias

101

Provide definitions or explain the meaning of the following terms: forward-bias voltage forward voltage fuse Grapher i–υ response ideal diode impede in series Interactive Simulation Kirchhoff’s current law (KCL) Kirchhoff’s voltage law (KVL) knee voltage law of conservation of charge law of conservation of energy light-emitting diode linear region linear resistor

resistive circuit resistivity reverse bias rheostat schematic capture window semiconductor siemen source transformation superconductor SPICE thermistor variable resistance voltage divider Wye–Delta (Y–) transformation

mechanical stress Multisim n-type negative NI myDAQ offset voltage Ohm’s law one-way valve overcurrent p-type piezoresistive coefficient piezoresistor pn-junction diode positive potentiometer power rating resistance

PROBLEMS

z 2 mm

Section 2-1: Ohm’s Law *2.1 An AWG-14 copper wire has a resistance of 17.1  at 20 ◦ C. How long is it? 2.2 A 3 km long AWG-6 metallic wire has a resistance of approximately 6  at 20 ◦ C. What material is it made of? 2.3 A thin-film resistor made of germanium is 2 mm in length and its rectangular cross section is 0.2 mm × 1 mm, as shown in Fig. P2.3. Determine the resistance that an ohmmeter would measure if connected across its: (a) Top and bottom surfaces

y 0.2 mm

1 mm x

Figure P2.3: Film resistor of Problem 2.3.

Carbon

l

Hollow

2a 2b

*(b) Front and back surfaces (c) Right and left surfaces 2.4 A resistor of length  consists of a hollow cylinder of radius a surrounded by a layer of carbon that extends from r = a to r = b, as shown in Fig. P2.4. (a) Develop an expression for the resistance R. (b) Calculate R at 20 ◦ C for a = 2 cm, b = 3 cm and  = 10 cm. ∗

Answer(s) available in Appendix G.

Figure P2.4: Carbon resistor for Problem 2.4.

2.5 A standard model used to describe the variation of resistance with temperature T is given by R = R0 (1 + αT ), where R is the resistance at temperature T (measured in ◦ C), R0 is the resistance at T = 0 ◦ C, and α is a

“book” — 2015/5/4 — 6:57 — page 102 — #53

102

CHAPTER 2

temperature coefficient. For copper, α = 4 × 10−3 ◦ C−1 . At what temperature is the resistance greater than R0 by 1 percent? 2.6 A light bulb has a filament whose resistance is characterized by a temperature coefficient α = 6 × 10−3 ◦ C−1 (see resistance model given in Problem 2.5). The bulb is connected to a 100 V household voltage source via a switch. After turning on the switch, the temperature of the filament increases rapidly from the initial room temperature of 20 ◦ C to an operating temperature of 1800 ◦ C. When it reaches its operating temperature, it consumes 80 W of power.

Ia 80 mA

4 cm 10 cm

7.5 cm

10 cm

2.5 cm 6 cm Ib

(a) Determine the filament resistance at 1800 ◦ C. (b) Determine the filament resistance at room temperature. (c) Determine the current that the filament draws at room temperature and also at 1800 ◦ C. (d) If the filament deteriorates when the current through it approaches 10A, is the damage done to the filament greater when it is first turned on or later when it arrives at its operating temperature? *2.7 A 110 V heating element in a stove can boil a standardsize pot of water in 1.2 minutes, consuming a total of 136 kJ of energy. Determine the resistance of the heating element and the current flowing through it.

RESISTIVE CIRCUITS

Figure P2.9: Circuit for Problem 2.9.



12 V

2.8 A certain copper wire has a resistance R characterized by the model given in Problem 2.5 with α = 4 × 10−3 ◦ C−1 . If R = 60  at 20 ◦ C and the wire is used in a circuit that cannot tolerate an increase in the magnitude of R by more than 10 percent over its value at 20 ◦ C, what would be the highest temperature at which the circuit can be operated within its tolerance limits?



+

+ _



10 Ω









Figure P2.10: Circuit for Problem 2.10.

Section 2-2: Kirchhoff’s Laws

2.10

I0

12 V _ +

2.9 The circuit shown in Fig. P2.9 includes two identical potentiometers with per-length resistance of 20 /cm. Determine Ia and Ib .

3I0

Determine VL in the circuit of Fig. P2.10.

*2.11 Select the value of R in the circuit of Fig. P2.11 so that VL = 9 V. 2.12 A high-voltage direct-current generating station delivers 10 MW of power at 250 kV to a city, as depicted in Fig. P2.12. The city is represented by resistance RL and each of the two wires of the transmission line between the generating station and the city is represented by resistance RTL . The distance between the two locations is 2000 km and the transmission lines are made of 10 cm diameter copper wire. Determine (a) how much power is consumed by the transmission line and (b)

R

500 Ω 6 mA

_ VL

+

500 Ω Figure P2.11: Circuit for Problem 2.11.

V

_L

“book” — 2015/5/4 — 6:57 — page 103 — #54

PROBLEMS

103

what fraction of the power generated by the generating station is used by the city.

*2.15

RTL V0

+ _

Determine Ix in the circuit of Fig. P2.15.

12 V RL

(city)

Ix



+ + __

1A



Figure P2.15: Circuit for Problem 2.15.

RTL

Station 2000 km

2.16

Determine currents I1 to I4 in the circuit of Fig. P2.16.

Figure P2.12: Diagram for Problem 2.12.

4A

1Ω 2.13 Determine the current I in the circuit of Fig. P2.13 given that I0 = 0.

12 V

I1

+ + __

I2



24 V

I



1Ω 2Ω

+ + __

5V

I4

Figure P2.16: Circuit for Problem 2.16.

I0 = 0





Figure P2.13: Circuit for Problem 2.13.

2.14

+ + __

6Ω 1V +_ +_



I3



*2.17

Determine currents I1 to I4 in the circuit of Fig. P2.17.

I1

I2





6A

I3

I4





Determine currents I1 to I3 in the circuit of Fig. P2.14. Figure P2.17: Circuit for Problem 2.17.

1A 2Ω

I1 18 V

+ + __

2.18 Determine the amount of power dissipated in the 3 k resistor in the circuit of Fig. P2.18.

3A I2 12 Ω

8Ω 4Ω

I3

Figure P2.14: Circuit for Problem 2.14.

7Ω 10 mA

+ V0 _

2 kΩ

3 kΩ

Figure P2.18: Circuit for Problem 2.18.

10−3V0

“book” — 2015/5/4 — 6:57 — page 104 — #55

104

CHAPTER 2

*2.19

Determine Ix and Iy in the circuit of Fig. P2.19.

2Ω 10 V

Ix



+ + __

Iy

+ V1 _

0.2 A

RESISTIVE CIRCUITS



V1 4

2Ω _ +



4Ix Figure P2.23: Circuit for Problem 2.23.

Figure P2.19: Circuit for Problem 2.19.

2.20

2.24 Given that in the circuit of Fig. P2.24, I1 = 4 A, I2 = 1A, and I3 = 1A, determine node voltages V1 , V2 , and V3 .

Find Vab in the circuit of Fig. P2.20.

2Ω a

2Ω + _

+ Vab _

6V

I2

2Ω 12 V

+ _

b

I1

1 Ω V1

R1 = 18 Ω 6Ω

+ 40 V _

V2



V3



18 Ω I3

Figure P2.20: Circuit for Problem 2.20. Figure P2.24: Circuit for Problem 2.24.

Find I1 to I3 in the circuit of Fig. P2.21.

3 kΩ I1 + + __ 16 V

I3 I2 4 kΩ

*2.25 After assigning node V4 in the circuit of Fig. P2.25 as the ground node, determine node voltages V1 , V2 , and V3 .

8V __ + +

2.21

2 kΩ

+ + __ 12 V 3A

12 V _ +

Figure P2.21: Circuit for Problem 2.21.



V1

V2

6Ω 2.22



Find I in the circuit of Fig. P2.22.

1A I 10 V

+ + __

2I



V4

V3 6Ω

1A

Figure P2.25: Circuit for Problems 2.25 and 2.26.

+_ 3Ω

Figure P2.22: Circuit for Problem 2.22.

2.26 After assigning node V1 in the circuit of Fig. P2.25 as the ground node, determine node voltages V2 , V3 , and V4 .

*2.23 Determine the amount of power supplied by the independent current source in the circuit of Fig. P2.23.

2.27 In the circuit of Fig. P2.27, I1 = 42/81 A, I2 = 42/81 A, and I3 = 24/81 A. Determine node voltages V2 , V3 , and V4 after assigning node V1 as the ground node.

“book” — 2015/5/4 — 6:57 — page 105 — #56

PROBLEMS

105

9Ω V2

6Ω I2

6Ω V1

6V __ + +





Find I0 in the circuit of Fig. P2.31.

V4

6V __



I1

+ +

V3

2.31

I3

I0 18 A

12 Ω

2.28 The independent source in Fig. P2.28 supplies 48 W of power. Determine I2 .

+ 12 V _

I3

2.32

Ix

t=0 1



+ + __ 15 V 2 Ω

0.25I1

R



For the circuit in Fig. P2.32, find Ix for t < 0 and t > 0.

R

I2

R



Figure P2.31: Circuit for Problem 2.31.

Figure P2.27: Circuit for Problem 2.27.

I1





2







R Figure P2.32: Circuit with SPDT switch for Problem 2.32.

Figure P2.28: Circuit for Problem 2.28.

Section 2-3: Equivalent Circuits *2.29 Given that I1 = 1 A in the circuit of Fig. P2.29, determine I0 .

2.33 Determine Req at terminals (a, b) in the circuit of Fig. P2.33.

I1 = 1 A I0









a

Req

16 Ω



32 Ω 16 Ω

8Ω 8Ω

b

Figure P2.29: Circuit for Problem 2.29.

Figure P2.33: Circuit for Problem 2.33.

2.30 What should R be in the circuit of Fig. P2.30 so that Req = 4 ?

a Req b

*2.34

Select R in the circuit of Fig. P2.34 so that VL = 5 V.

1Ω 6Ω



R

5 mA

R 5 kΩ

1 kΩ 2 kΩ

5Ω Figure P2.30: Circuit for Problem 2.30.

Figure P2.34: Circuit for Problem 2.34.

+ VL _

“book” — 2015/5/4 — 6:57 — page 106 — #57

106

CHAPTER 2 If R = 12  in the circuit of Fig. P2.35, find I .

2.35

R

2.39

Find Req at terminals (c, d) in the circuit of Fig. P2.38.

2.40 Simplify the circuit to the right of terminals (a, b) in Fig. P2.40 to find Req , and then determine the amount of power supplied by the voltage source. All resistances are in ohms.

R R

RESISTIVE CIRCUITS

R 20 V __



a

I

+ +

25 V R

+ + __

Req

3

5 8

4

8

6

12

6

12

R

R

R

b Figure P2.40: Circuit for Problem 2.40.

Figure P2.35: Circuit for Problem 2.35.

2.41 *2.36 Use resistance reduction and source transformation to find Vx in the circuit of Fig. P2.36. All resistance values are in ohms.

16

16

10 A

12

4

6

*(a) Terminals (a, b) (b) Terminals (a, c) (c) Terminals (a, d) (d) Terminals (a, f )

+ Vx _

4

For the circuit in Fig. P2.41, determine Req at

16

16 e

Figure P2.36: Circuit for Problem 2.36.

2.37

d

Determine A if Vout /Vs = 9 in the circuit of Fig. P2.37.









c

3Ω Vs

I1

+ + __

12 Ω

12 Ω



AI1

+ 6 Ω Vout _

b









a

b

Figure P2.41: Circuit for Problem 2.41.

For the circuit in Fig. P2.38, find Req at terminals (a, b).

5 a





2.42 Find Req for the circuit in Fig. P2.42. All resistances are in ohms.

Figure P2.37: Circuit for Problem 2.37.

*2.38

f



3Ω 6Ω



5Ω 6Ω



Figure P2.38: Circuit for Problems 2.38 and 2.39.

c d

Req

10 10

5

10

10 10

Figure P2.42: Circuit for Problem 2.42.

“book” — 2015/5/4 — 6:57 — page 107 — #58

PROBLEMS

107

2.43 Apply voltage and current division to determine V0 in the circuit of Fig. P2.43 given that Vout = 0.2 V.

3A

6Ω 4Ω





3Ω I

4Ω + _ V0

+_

2A

30 V





2Ω 2Ω 1Ω

+ Vout = 0.2 V _

Figure P2.46: Circuit for Problem 2.46.

2.47

Figure P2.43: Circuit for Problem 2.43.

Determine currents I1 to I4 in the circuit of Fig. P2.47.

12 Ω

I1



I2

I3



I4



*2.44 Apply source transformations and resistance reductions to simplify the circuit to the left of nodes (a, b) in Fig. P2.44 into a single voltage source and a resistor. Then, determine I .

3A

10 Ω

+ _

12 V

a I

5A

12 Ω



Figure P2.47: Circuit for Problems 2.47 and 2.48.



2.48 Replace the 12 V source in the circuit of Fig. P2.47 with a 4 A current source pointing upwards. Then, determine currents I1 to I4 .

b Figure P2.44: Circuit for Problem 2.44.

*2.49

Determine current I in the circuit of Fig. P2.49.

2.45 Determine the open-circuit voltage Voc across terminals (a, b) in Fig. P2.45.

10 Ω

40 Ω

25 Ω I

30 V

+ _

5Ω 3Ω

6Ω 2A

+

a

Voc

_

b

Figure P2.45: Circuit for Problem 2.45.

2.46 Use circuit transformations to determine I in the circuit of Fig. P2.46.

5Ω 30 Ω

60 Ω + _

50 V

10 Ω

Figure P2.49: Circuit for Problem 2.49.

10 Ω

“book” — 2015/5/4 — 6:57 — page 108 — #59

108

CHAPTER 2

RESISTIVE CIRCUITS

2.50 Determine the equivalent resistance Req at terminals (a, b) in the circuit of Fig. P2.50.



2A

_

V

a

+

2A



5Ω a





Req



2.5 A

5A 4Ω



4Ω b

Figure P2.52: Circuit for Problem 2.52.

6Ω Sections 2-4 and 2-5: Y– and Wheatstone Bridge Figure P2.50: Circuit for Problem 2.50.

2.53 Convert the circuit in Fig. P2.53(a) from a  to a Y configuration.

Determine current I in the circuit of Fig. P2.51.

*2.51

a 3Ω

+ _



d

b





d

2 kΩ

(b)

Figure P2.53: Circuit for Problems 2.53 and 2.54.

6 mA 5 mA

2 mA

2.54 Convert the circuit in Fig. P2.53(b) from a T to a

configuration. *2.55

Find the power supplied by the generator in Fig. P2.55.

R1 = 18 Ω

1 kΩ I

1Ω _ +

2.52

c

(a) 2 kΩ

2 kΩ

a



b

16 V

c



8V

+ 20 V _



6Ω 6Ω

18 Ω

Figure P2.51: Circuit for Problem 2.51.

Figure P2.55: Circuit for Problems 2.55 and 2.56.

Determine voltage Va in the circuit of Fig. P2.52.

2.56 Repeat Problem 2.55 after replacing R1 with a short circuit.

“book” — 2015/5/4 — 6:57 — page 109 — #60

PROBLEMS 2.57

109

Find I in the circuit of Fig. P2.57.

2.62 Find Req at terminals (a, b) in Fig. P2.62 if (a) Terminal c is connected to terminal d by a short circuit (b) Terminal e is connected to terminal f by a short circuit (c) Terminal c is connected to terminal e by a short circuit All resistance values are in ohms.

9Ω 6Ω

6Ω 9Ω



3V __

3V __



I

d

+ +

+ +

3 Figure P2.57: Circuit for Problem 2.57.

e 2.58 Find the power supplied by the voltage source in Fig. P2.58.

4V _

3

3 Req

3

a b 3

f

3 c

+ Figure P2.62: Circuit for Problem 2.62.



3Ω 6Ω



R=6Ω

Figure P2.58: Circuit for Problems 2.58 and 2.59.

*2.59 Repeat Problem 2.58 after replacing R with a short circuit. 2.60 Find I in the circuit of Fig. P2.60. All resistances are in ohms.

I

1 12 V

2

+ + __

4

2

2 2

Figure P2.60: Circuit for Problem 2.60.

*2.61

Find Req for the circuit in Fig. P2.61.

18 Ω 6Ω







18 Ω

Figure P2.61: Circuit for Problem 2.61.

2.64 If V0 = 10 V in the Wheatstone-bridge circuit of Fig. 2-37 and the minimum voltage Vout that a voltmeter can read is 1 mV, what is the smallest resistance fraction (R/R) that can be measured by the circuit? 2.65 Suppose the cantilever system shown in Fig. 2-44 is used in the Wheatstone-bridge sensor of Fig. 2-37 with V0 = 2 V, α = −1 × 10−9 m2 /N, L = 0.5 cm, W = 0.2 cm, and H = 0.2 mm. If the measured voltage is Vout = −2 V, what is the force applied to the cantilever? *2.66 A touch sensor based on a piezoresistor built into a micromechanical cantilever made of silicon is connected in a Wheatstone-bridge configuration with a V0 = 1 V. If L = 1.44 cm and W = 1 cm, what should the thickness H be so that the touch sensor registers a voltage magnitude of 10 mV when the touch pressure is 10 N?

18 Ω 6Ω

2.63 For the Wheatstone-bridge circuit of Fig. 2-36, solve the following problems. *(a) If R1 = 1 , R2 = 2 , and Rx = 3 , to what value should R3 be adjusted so as to achieve a balanced condition? (b) If V0 = 6 V, Ra = 0.1 , and Rx were then to deviate by a small amount to Rx = 3.01 , what would be the reading on the ammeter?

Req

Section 2-6: i–υ Relationships *2.67 Determine I1 and I2 in the circuit of Fig. P2.67. Assume VF = 0.7 V for both diodes.

“book” — 2015/5/4 — 6:57 — page 110 — #61

110

CHAPTER 2

53 Ω 6V

I1

I2

53 Ω

73 Ω

i1

RESISTIVE CIRCUITS

i2

146 Ω

+ _ υs(t)

+ _

Figure P2.67: Circuit for Problem 2.67.

2.68 Determine V1 in the circuit of Fig. P2.68. Assume VF = 0.7 V for all diodes.

(a)

υs(t)

50 Ω

9V

+

100 Ω

+ _

8V

V_1

25 Ω

t (s)

4

2 Figure P2.68: Circuit for Problem 2.68.

−8 V 2.69 If the voltage source in the circuit of Fig. P2.69 generates a single square wave with an amplitude of 2 V, generate a plot for vout for the same time period.

(b) Square wave υs(t)

υs(t)

+ _

100 Ω

+ _

8V

υout

2 υs(t)

4

t (s)

2V T

t

−8 V (c) Triangular wave

−2 V

Figure P2.70: Circuit and waveforms for Problems 2.70 and Figure P2.69: Circuit and voltage waveform for Problem 2.69.

2.70 If the voltage source in the circuit of Fig. P2.70(a) generates the single square waveform shown in Fig. P2.70(b), generate plots for i1 (t) and i2 (t).

2.71.

2.71 If the voltage source in the circuit of Fig. P2.70(a) generates the single triangular waveform shown in Fig. P2.70(c), generate plots for i1 (t) and i2 (t).

“book” — 2015/5/4 — 6:57 — page 111 — #62

PROBLEMS

111

2.72 The circuit shown in Fig. P2.72 is used to control a red LED. The LED is designed to turn on when the resistance R of the rheostat is 50  or lower. Use the information contained in Fig. 2-8(d) to determine the value of the constant resistor R0 .

2.75 Use DC Operating Point Analysis in Multisim to solve for all six labeled resistor currents in the circuit of Fig. P2.75.



I R

R0

+ 5V _

I2

I1

+ _VF

1Ω +_

I3 1A

Red LED





RD

+_

I5



Figure P2.75: Circuit for Problem 2.75.

Section 2-7: Multisim 2.73 Use the DC Operating Point Analysis in Multisim to solve for voltage Vout in the circuit of Fig. P2.73. Solve for Vout by hand and compare with the value generated by Multisim. See the solution for Exercise 2.14 (on ) for how to incorporate circuit variables into algebraic expressions.

10 Ω

+ 2.5 V _

I6

3V



Figure P2.72: Circuit for Problem 2.72.

I4

2V

10 Ω

+

2.76 Find the voltages across R1 , R2 , and R3 in the circuit of Fig. P2.76 using the DC Operating Point Analysis tool in Multisim.

R1 V1

Vout

10 Ω + _ 15 V

R3 I R2

15 Ω

30 Ω

1.5I

_ 15 Ω

Figure P2.76: Circuit for Problem 2.76.

25 Ω

Figure P2.73: Circuit for Problem 2.73.

2.74 Find the ratio Vout /Vin for the circuit in Fig. P2.74 using DC Operating Point Analysis in Multisim. See the Multisim Tutorial included on the book website on how to reference currents in ABM sources [you should not just type in I(V1)].

2.77 Find the equivalent resistance looking into the terminals of the circuit in Fig. P2.77 using a test voltage source and current probes in the Interactive Simulation in Multisim. Compare the answer you get to what you obtain from series and parallel combining of resistors carried out by hand. Potpourri Questions

1 kΩ Iin + Vin _

10 kΩ

+ 100Iin

Vout

_

Figure P2.74: Circuit for Problem 2.74.

1 kΩ

2.78 What is a superconducting material and what happens when its physical temperature is below or above its critical temperature? How is superconductivity used in practice? 2.79 What is a piezoresistor? How is it used? Resistors are also used as chemical sensors. Explain how. 2.80 What determines the color of the light emitted by an LED? Why are LEDs economical to use?

“book” — 2015/5/4 — 6:57 — page 112 — #63

112

CHAPTER 2

RESISTIVE CIRCUITS

Figure P2.77: Circuit for Problem 2.77.

Integrative Problems: Analytical / Multisim / myDAQ

(a) a-b with the other terminals unconnected,

To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available in Appendix F and on .]

(b) a-d with the other terminals unconnected,

m2.1 Kirchhoff’s Laws: Determine currents I1 to I3 and the voltage V1 in the circuit of Fig. m2.1 with component values Isrc = 1.8 mA, Vsrc = 9.0 V, R1 = 2.2 k, R2 = 3.3 k, and R3 = 1.0 k.

Vsrc

R1

I1

R2

I2

(d) a-d with a wire connecting terminals b and c. Use these component values: R1 = 10 k, R3 = 15 k, R4 = 47 k, and R5 = 22 k.

a

R1

R2

R3

R2 = 33 k,

b

R5

+ _ c

_ Isrc

(c) b-c with a wire connecting terminals a and d, and

V1

+

R4

d

Figure m2.2 Circuit for Problem m2.2.

R3

I3

Figure m2.1 Circuit for Problem m2.1.

m2.2 Equivalent Resistance: Find the equivalent resistance between the following terminal pairs in the circuit of Fig. m2.2 under the stated conditions:

m2.3 Current and Voltage Dividers: Apply the concepts of voltage dividers, current dividers, and equivalent resistance to find the currents I1 to I3 and the voltages V1 to V3 in the circuit of Fig. m2.3. Use these component values: Vsrc = 12 V, R1 = 1.0 k, R2 = 10 k, R3 = 1.5 k, R4 = 2.2 k, R5 = 4.7 k, and R6 = 3.3 k.

“book” — 2015/5/4 — 6:57 — page 113 — #64

PROBLEMS

113

R4

I1

R1

R3

33 kΩ

+ _

R3

Vsrc

I2

+

+ _

V1

_

R2

+ V3

R1 1 kΩ

46 kΩ R2 2.2 kΩ R5

R6

100 kΩ

3.3 kΩ

+ V2 R4

R5

_

R6

V1 15 V

Figure m2.5 Circuit for Problem m2.5.

I3 m2.6 Multiple Sources: To create multiple sources,use the AO 0 and AO 1 ports simultaneously for the myDAQ portion of this problem. Use the Arbitrary waveform generator to create the 3 V and 5 V sources.

_

Figure m2.3 Circuit for Problem m2.3.

m2.4 Wye-Delta Transformation: Find (a) the currents I1 and I2 in the circuit of Fig. m2.4 and (b) the power delivered by each of the two voltage sources. Use these component values: V1 = 15 V, V2 = 15 V, R1 = 3.3 k, R2 = 1.5 k, R3 = 4.7 k, R4 = 5.6 k, R5 = 1.0 k, and R6 = 2.2 k.

(a) Find currents I1 and I2 in the circuit of Fig. m2.6. For the myDAQ portion, make sure to measure current correctly or you could blow the myDAQ’s fuse. (b) Find the voltage drop across the 47 k resistor.

R2 47 kΩ R5

R3

R3 10 kΩ

R2

R1

R5

I1

+ _ R4 V1 _

+

I1

R6 V2 _

I2

22 kΩ V1 3V

+ _

V2 5V

R4

10 kΩ

Figure m2.6 Circuit for Problem m2.6.

+

R1

1 kΩ

I2 Figure m2.4 Circuit for Problem m2.4.

m2.5 Kirchoff’s Laws and Equivalent Resistance: In the circuit of Fig. m2.5:

m2.7 Current Source: This problem is relatively straightforward to solve by hand and with Multisim. However, to create the myDAQ version of the circuit in Fig. m2.7, you will need to use an LM371 Regulator with a 100  connected between Vout and Vadj . For more information, consult Appendix F or look up the specification of the LM371-LZ regulator.

(a) Find the voltage drop across the 46 k resistor.

(a) Determine the voltage drop across each 1 k resistor.

(b) What is the equivalent resistance seen by the 15 V source?

(b) Determine the current through the 3.3 k resistor.

“book” — 2015/5/4 — 6:57 — page 114 — #65

114

CHAPTER 2

m2.8 Equivalent Resistance: Determine the equivalent resistance of the circuit in Fig. m2.8 as seen at terminals (1, 2).

R2 I1

12.5 mA

R1

3.3 kΩ 1 kΩ

RESISTIVE CIRCUITS

R3

1 kΩ R1 33 kΩ

Figure m2.7 Circuit for Problem m2.7.

R2 1 kΩ

1 R3

46 kΩ R5 1 kΩ R4

R6 10 kΩ

2.2 kΩ

Figure m2.8 Circuit for Problem m2.8.

2

“book” — 2015/5/4 — 7:08 — page 115 — #1

3

CHAPTER

Analysis Techniques Contents 3-1 3-2 3-3 TB6 3-4 3-5 TB7 3-6 3-7 3-8 TB8 3-9 3-10

Overview, 116 Linear Circuits, 116 Node-Voltage Method, 117 Mesh-Current Method, 123 Measurement of Electrical Properties of Sea Ice, 126 By-Inspection Methods, 129 Linear Circuits and Source Superposition, 133 Integrated Circuit Fabrication Process, 136 Th´evenin and Norton Equivalent Circuits, 140 Comparison of Analysis Methods, 151 Maximum Power Transfer, 151 Digital and Analog, 154 Application Note: Bipolar Junction Transistor (BJT), 158 Nodal Analysis with Multisim, 161 Summary, 164 Problems, 165

RB

B IB

C

+

IC VBB

VBE

βIB

RC

VCC

VCE

_

E Transistor equivalent circuit The basic laws of Chapter 2 are used in the present chapter to develop standard solution methods that can be applied to analyze any linear circuit, no matter how complex.

Objectives Learn to: 

Apply the node-voltage and mesh-current methods to analyze an electric circuit of any configuration, so long as it is linear and planar.



Apply the by-inspection methods to circuits that satisfy certain conditions.



Use the source-superposition method to evaluate the sensitivity of a circuit to the various sources in the circuit.



Determine the Th´evenin and Norton equivalent circuits of any input circuit and use them to evaluate the response of an external load (or an output circuit) to the input circuit.



Establish the conditions for maximum transfer of current, voltage, and power from an input circuit to an external load.



Learn the basic properties of the bipolar junction transistor.

“book” — 2015/5/4 — 7:08 — page 116 — #2

116

CHAPTER 3 ANALYSIS TECHNIQUES

Overview

3-1.2

By applying the circuit-analysis skills we developed in the preceding chapter, we now extend our capability further so we may tackle any linear, planar circuit—no matter how complex. Node-voltage and mesh-current equations will be cast into a systematic structure in Sections 3-2 through 3-4, so we may take advantage of standard methods for solving linear, simultaneous equations, either by the use of determinants and matrices (Appendix B) or the execution of computer simulation packages such as MATLAB or MathScript (Appendix E). The nodal and mesh analysis techniques are followed with treatments of two special tools: the source superposition method and the Th´evenin/Norton equivalent-circuit method. These methods allow us to break any complex electrical system into smaller, manageable subcircuits for analysis. With these tools, you are ready to analyze pretty much any circuit you may encounter for the rest of your career. We will also introduce you to semiconductor manufacturing and the relationships between analog and digital signals.

If current i1 can give rise to voltage υ1 = Ri1 , and another current i2 can give rise to voltage υ2 = Ri2 , then the simultaneous presence of both currents gives rise to

Superposition Principle

υ = R(i1 + i2 ) = Ri1 + Ri2 = υ1 + υ2 .

(3.2)

Thus, the output (υ) due to the two inputs (i1 and i2 ) is equal to the sum of the two outputs (υ1 and υ2 ) had each input been introduced separately. This is a statement of the superposition principle (also known as the additivity property). We will use this principle in Section 3-5 to simplify our analysis for circuits containing multiple sources.

3-1.3

Linear and Nonlinear Elements

Linear elements

3-1 Linear Circuits

By virtue of its linear i–υ relationship, the resistor is an obvious candidate for the list of linear circuit elements, which includes:

A circuit is a system with inputs and outputs; its inputs are the independent voltage and current sources that energize the circuit, and its outputs are all of the currents flowing through and voltages across all of the passive elements of the circuit. By passive element, we mean that it does not generate energy of its own. A resistor is a perfect example of a passive element. By comparison, an active element requires an external power supply in order to function. Examples of active elements include transistors (such as the BJT described in Section 3-9) and operational amplifiers (Chapter 4). A linear circuit is a circuit composed entirely of independent sources and linear elements. An element is linear if it is passive and exhibits a linear i–υ relationship. For a resistor R, for example, υ = Ri.

(3.1)

A circuit element, or an entire circuit, is nonlinear if its i–υ relationship is not linear. The LED (Section 2-1.4) is an example of a nonlinear device.

3-1.1

Homogeneity Property

If i through resistor R is increased by a factor K, so will υ. This proportional increase of i and υ by the same factor is called the homogeneity (or scaling) property of a linear element.

• Resistors • Capacitors • Inductors • Linear dependent sources The i–υ relationship for a capacitor, which we will learn more about in Chapter 5, is given by i=C

dυ . dt

(3.3a)

If we multiply both sides by a factor K, we get Ki = KC

dυ d =C (Kυ). dt dt

(3.3b)

Hence, increasing υ by a factor K leads to an increase in i by the same factor, which implies that the d/dt differentiation operator has no bearing on the homogeneity property linking i to υ. The time derivative does not impact the additivity property either.

“book” — 2015/5/4 — 7:08 — page 117 — #3

3-2

NODE-VOLTAGE METHOD

+ Vs _

I1

R1 R2

117

R3 + _

Va = 5I1

Figure 3-1: Circuit with dependent source Va = 5I1 .

(2) However, it is often possible to replace nonlinear elements with equivalent circuits containing linear elements, including dependent sources, and then use them to obtain approximate, but fairly accurate results, provided certain conditions are satisfied. Examples of equivalent circuits will be presented in Section 3-8 for the bipolar junction transistor (BJT) and in Chapter 4 for the operational amplifier and the CMOS transistor.

3-1.4 Advantages of Linear Circuits

 Since the capacitor is a passive element and obeys both the homogeneity and additivity (superposition) properties, it is classified as a linear circuit element. A similar argument applies to the inductor, for which υ = L di/dt.  Next we consider dependent sources, which were first introduced in Section 1-6.4. Dependent sources are artificial sources (because they do not generate energy of their own) used in equivalent linear circuits intended to model the approximate behavior of nonlinear circuits and elements like transistors and operational amplifiers. Let us consider the simple circuit shown in Fig. 3-1, which includes an independent voltage source Vs and a dependent voltage source Va . The magnitude of Va depends on I1 , which, in turn, depends on the real source Vs . If Vs = 0, no currents would flow in the circuit, so I1 would be zero, and so would Va .  Hence, dependent source Va is a passive element, and since it is also directly proportional to I1 (raised to first order), Va is classified as a linear element. The same is true for a dependent voltage source whose magnitude is linearly related to a voltage elsewhere in the circuit (instead of to a current), as well as for dependent current sources that depend linearly on a voltage or current elsewhere in the circuit.  Nonlinear elements The circuit analysis techniques developed in this book apply primarily to linear circuits, and yet many devices—such as diodes, transistors, and integrated circuits—exhibit nonlinear i–υ relationships. Consequently: (1) The analysis techniques do not directly apply to circuits containing such nonlinear elements.

The linearity properties of a linear circuit allow us to use certain analysis techniques that would be otherwise not applicable had the circuit contained one or more nonlinear elements (unless they can be adequately represented by equivalent linear circuits). Through the application of such analysis techniques, which include the Th´evenin and superposition methods presented later in Sections 3-5 and 3-6, we can simplify the analysis (and design) of a complex circuit considerably.

3-2

Node-Voltage Method

3-2.1

General Procedure

According to Kirchhoff’s current law (KCL), the algebraic sum of all currents entering any node in an electric circuit is equal to zero. Built on that principle, the node-voltage analysis method provides a systematic and efficient procedure for determining all of the currents and voltages in a circuit. This determination is realized through the solution of a system of linear, simultaneous equations in which the unknown variables are the voltages at the extraordinary nodes in the circuit. As a reminder, in Section 1-3 we defined an extraordinary node as a node connected to three or more elements. For a circuit containing nex extraordinary nodes, implementation of the node-voltage method consists of three basic steps:

Solution Procedure: Node Voltage Step 1: Identify all extraordinary nodes, select one of them as a reference node (ground), and then assign node voltages to the remaining (nex − 1) extraordinary nodes. Step 2: At each of the (nex − 1) extraordinary nodes, apply the form of KCL requiring the sum of all currents leaving a node to be zero (see KCL template). Step 3: Solve the (nex − 1) independent simultaneous equations to determine the unknown node voltages (see Appendix B).

“book” — 2015/5/4 — 7:08 — page 118 — #4

118

CHAPTER 3 ANALYSIS TECHNIQUES

KCL Template V1

R1

I1

V0

I3

R3

I2 R2

R2

V3

R3

+ _

V0

R4 R1

R5

I0

I0

V2

(a) υ = V0

V0 − V 1 V0 − V2 V0 − V3 + + − I0 = 0 R1 R2 R3 Once the node voltages have been determined, all currents through branches and voltages across elements can be calculated readily.

Original circuit I2 R2 + R3

+ _

V0

For the circuit in Fig. 3-2, (a) identify all extraordinary nodes and select one of them as the ground node, (b) develop node-voltage equations at the remaining extraordinary nodes, (c) solve for the node voltages, and then (d) calculate the power consumed by R5 . The element values are V0 = 10 V, I0 = 0.8 A, R1 = 5 , R2 = 2 , R3 = 3 , R4 = 10 , and R5 = 2.5 . Solution: (a) Identify extraordinary nodes and assign node voltages The circuit has three extraordinary nodes, labeled as shown in Fig. 3-2(b). Node 3 is selected as the ground node and its voltage is labeled V3 = 0. Nodes 1 and 2 are assigned (unknown) voltages V1 and V2 , with both defined relative to V3 = 0. (b) Apply KCL at nodes 1 and 2 At each non-ground extraordinary node, we designate currents and we choose their directions as leaving the node. We realize that I3 = −I4 , for example, but for the sake of consistency we treat each node the same by designating a current leaving it through every branch connected to it.

(b)

I4

I6

V2

R4

I1

I5

R1

R5

V3 = 0

υ=0 Example 3-1: Circuit with Two Sources

I3

V1

I0

υ=0

Circuit with designated node voltages Figure 3-2: Circuit for Example 3-1.

Node 1: I1 + I2 + I3 = 0.

(3.4)

Unless we already know the value of a current (such as I0 entering node V2 ), we should express it in terms of the node voltages connected to the branch through which it is flowing. We do so by applying Ohm’s law, while reminding ourselves that the convention we adopted for the current direction is that it flows through a resistor from the (+) voltage terminal to the (−) terminal. Hence:  The current leaving a node is equal to the voltage at that node, minus the voltage at the node to which the current is going, and divided by the resistance.

υa

R

+ i=

υa − υb R

_

υb 

“book” — 2015/5/4 — 7:08 — page 119 — #5

3-2

NODE-VOLTAGE METHOD

119

Consequently, I1 flowing through R1 is given by I1 =

V1 − 0 V1 = . R1 R1

(c) Solve simultaneous equations (3.5a)

As a prelude to solving Eqs. (3.6) and (3.7) to determine the unknown voltages V1 to V3 , we need to reorganize them into a standard system of equations as 

Similarly, V1 − V2 . I3 = R4

(3.5b)

The voltage across the in-series resistances (R2 + R3 ) is (V1 − V0 ), where V0 is the node voltage at the positive terminal of the voltage source. Hence, I2 is given by I2 =

V 1 − V0 . R2 + R 3

(3.5c)

Inserting Eqs. (3.5a) through (3.5c) into Eq. (3.4) gives V1 − V0 V1 − V2 V1 + + =0 R1 R2 + R 3 R4

1 1 1 + + R1 R2 + R 3 R4

and

 −



 V1 −

 V1 +

1 R4

1 1 + R4 R5

 V2 =

V0 , R2 + R 3 (3.8a)

 V2 = I0 .

(3.8b)

These are equivalent to a11 V1 + a12 V2 = b1 ,

(3.9a)

a21 V1 + a22 V2 = b2 ,

(3.9b)

and

(node 1 Voltage Eq.). (3.6)

with  a11 =

Node 2: I4 + I5 + I6 = 0, or equivalently, V2 V2 − V1 + − I0 = 0 R4 R5

1 R4



where we incorporated the fact that I6 = −I0 , as required by the current source. We note that by designating all current directions at a node as leaving that node:  The node-voltage expression for any node (such as node 1 or node 2) always has V of that node preceded with a plus (+) sign. Also, the node voltages of the other nodes are preceded with negative (−) signs.  Thus, V1 in Eq. (3.6)—which is specific to node 1—has a positive sign wherever it appears in that equation, whereas V2 and V3 always have negative signs if they appear in that equation. Conversely, in the node-2 equation given by Eq. (3.7), V2 is always preceded by a (+) sign and V1 is preceded by a (−) sign.

 =

1 1 1 + + = 0.5, 5 2 + 3 10

1 1 =− = −0.1, R4 10 1 =− = −0.1, R4   1 1 1 1 = + = 0.5, = + R4 R5 10 2.5

a12 = − a21

(node 2 Voltage Eq.), (3.7)

1 1 1 + + R1 R2 + R 3 R4

a22

b1 =

V0 10 = = 2, R2 + R 3 2+3

and b2 = I0 = 0.8. Inserting these values in Eq. (3.9) gives 0.5V1 − 0.1V2 = 2, −0.1V1 + 0.5V2 = 0.8. The system of two equations is now amenable for solution by Cramer’s rule or matrix inversion (as illustrated in Appendix B) either manually or by using MATLAB or MathScript software (Appendix E). The solution leads to V1 = 4.5 V,

V2 = 2.5 V.

“book” — 2015/5/4 — 7:08 — page 120 — #6

120

CHAPTER 3 ANALYSIS TECHNIQUES

(d) Determine power in R5



The current flowing through R5 in Fig. 3-2(b) is

+ _

V2 2.5 = 1 A, = I5 = R5 2.5

6Ω I 3Ω

5.3 V

Ix = 2I

12 Ω

and the power dissipated in R5 is P = I52 R5 = (1)2 × 2.5 = 2.5 W.

(a) Original circuit

Concept Question 3-1: The node-voltage method relies

on the application of Kirchhoff’s current law. Explain. (See )

5.3 V 4 Ω + _

Concept Question 3-2: Why does a circuit with nex

5.3 V

I1

V1

I3



I2

I



I4 V I6 2 I5 12 Ω

Ix = 2I

extraordinary nodes require only (nex − 1) node-voltage equations to analyze it? (See ) Exercise 3-1: Apply nodal analysis to determine the current I in the circuit of Fig. E3.1.

(b) Circuit with designated node voltages

I



Figure 3-3: Example 3-2.

10 Ω 4Ω 24 V



+ _

Figure E3.1 Answer: I = 2 A. (See

3-2.2

)

flowing through the 6  resistor in the direction shown. Determine Ix . Solution: Following the standard procedure outlined earlier, we start by selecting a ground node and assigning node voltages to the other extraordinary nodes in the circuit, as shown in Fig. 3-3(b). We also designate currents with their directions out of the nodes for all branches connected to nodes 1 and 2. Next, we write down the node-voltage equations for nodes 1 and 2 as

Circuits Containing Dependent Sources

When a circuit contains dependent sources, the node-voltage analysis method remains applicable, as does the solution procedure outlined in the preceding subsection. However, each dependent source defines a relationship between its own magnitude and some current or voltage elsewhere in the circuit, and that relationship needs to be incorporated into the solution. Example 3-2: Dependent Current Source

The circuit of Fig. 3-3 contains a current-controlled current source (CCCS) whose magnitude Ix is governed by the current

V1 − 5.3 V1 V1 − V2 + + =0 4 3 6

(node 1),

V2 − V1 V2 + − Ix = 0 6 12

(node 2).

and

In the equation for node 1, the three terms represent I1 to I3 , each expressed as a voltage difference divided by a resistance. The same is true for node 2 except that I6 is replaced with (−Ix ). We have three unknowns (V1 , V2 , and Ix ), but only two equations, so we need to express Ix in terms of the unknown variables, V1 and V2 . The dependent source Ix is given in terms of I , which in turn is dependent on the voltage difference

“book” — 2015/5/4 — 7:08 — page 121 — #7

3-2

NODE-VOLTAGE METHOD

121

R1 Quasisupernode

Supernode A R2

V2

+ _

20 V

+

V1

10 V _

Supernode B R4

V3

R3

R5 R7

V4 + _ 16 V

R6

V5

Figure 3-4: Circuit containing two supernodes and one quasi-supernode.

between V1 and V2 . That is,

3-2.3

(V1 − V2 ) V1 − V2 Ix = 2I = 2 = . 6 3 This is effectively Ohm’s law for Ix . Upon substituting this expression for Ix into the second of the node-voltage equations and rearranging its terms, we end up with 9V1 − 2V2 = 15.9

(node 1)

Supernodes

Occasionally, a circuit may contain a solitary voltage source nestled between two extraordinary nodes, with no other elements in series with it between those nodes. Such an arrangement is called a supernode. Examples of supernodes are shown in Fig. 3-4. Formally:

and −6V1 + 7V2 = 0

(node 2).

Simultaneous solution of the two equations gives V1 = 2.18 V and V2 = 1.87 V. Hence, Ix =

V1 − V2 2.18 − 1.87 = = 0.1 A. 3 3

Exercise 3-2: Apply nodal analysis to find Va in the circuit

of Fig. E3.2.

 A supernode is the combination of two extraordinary nodes (excluding the reference node) between which a voltage source exists. The voltage source may be of the independent or dependent type, and the voltage source may include elements in parallel with it (such as R6 in parallel with the 16-V source of supernode B in Fig. 3-4) but not in series with it. If one of the two nodes of a supernode is a reference (ground) node, it is called a quasi-supernode. 

20 Ω + Va 9V

_ 10 Ω

+ _

Va 2

+ _

Figure E3.2 Answer: Va = 5 V. (See

)

40 Ω

For a quasi-supernode, the only relevant information we need is that the voltage of the non-reference node is equal to the voltage magnitude of the voltage source. Thus, V1 = 20 V in Fig. 3-4. The complication caused by a supernode is that we can no longer apply Ohm’s law to define the current through a resistor between two extraordinary nodes, because we now have a voltage source between the two nodes instead of a resistor. Hence, we need to treat the supernode in a special way. To explain the properties of a supernode and how we use it, let us analyze supernode A, all on its own. In Fig. 3-5(a),

“book” — 2015/5/4 — 7:08 — page 122 — #8

122

CHAPTER 3 ANALYSIS TECHNIQUES

I4

I2

Supernode

I6

V3

4V 2Ω

I5

R3

I1 V1 I2

R5

+ _

4V



18 V _+ − +

10 V _+ −

I3

V2

+

I1

I4

V2

I3 8Ω

2A

Supernode A (a)

I1

V2

Figure 3-6: Circuit for Example 3-3.

I6

V3 I2

I5

R3

R5

which is a much simpler equation than the typical node-voltage equation.

Supernode Attributes (1) At a supernode, Kirchhoff’s current law (KCL) can be applied to the combination of the two nodes as if they are a single node, but the two nodes retain their own identities.

I1 + I2 + I5 + I6 = 0 (KCL)

(b)

_

V3

+

V2

V3 − V2 = 10 V (KVL)

Figure 3-5: A supernode composed of nodes V2 and V3 can be represented as a single node, in terms of summing currents flowing out of them, plus an auxiliary equation that defines the voltage difference between V3 and V2 . we show currents I1 to I3 leaving node 2 and currents I4 to I6 leaving node 3. KCL requires that I 1 + I2 + I3 = 0

(node V2 ),

(3.10a)

I4 + I5 + I6 = 0

(node V3 ).

(3.10b)

and

Adding the two equations together and recognizing that I3 = −I4 leads to I1 + I2 + I5 + I6 = 0

(3) If a supernode contains a resistor in parallel with the voltage source, the resistor exercises no influence on the currents and voltages in the other parts of the circuit, and therefore, it may be ignored altogether. (4) For a quasi-supernode, the node-voltage of the nonreference node is equal to the voltage magnitude of the source. In the circuit of Fig. 3-4, the voltage difference between nodes 4 and 5 is specified by the 16 V source, regardless of the value of R6 (so long as R6 is not a short circuit). Example 3-3: Circuit with a Supernode

(supernode A),

(3.11)

which constitutes the four currents leaving supernode A. The implication of Eq. (3.11) is that we can treat nodes 2 and 3 as a combined single node, connected by a dashed line (Fig. 3-5(b)), but we also should acknowledge the fact that V3 − V2 = 10 V

(2) Kirchhoff’s voltage law (KVL) is used to express the voltage difference between the two nodes in terms of the voltage of the source between them. This provides the supernode auxiliary equation.

(supernode A auxiliary equation),

Use the supernode concept to solve for the node voltages in Fig. 3-6. Solution: The combination of nodes 1 and 2 constitutes a supernode, with an associated node-voltage equation given by I1 + I2 + I3 + I4 = 0

“book” — 2015/5/4 — 7:08 — page 123 — #9

3-3

MESH-CURRENT METHOD

123

or, equivalently,

R1

V1 − 4 V 1 V2 + + − 2 = 0, 2 4 8

V0

which may be simplified to

+ _

Ia

Ic

R2

Ib I1

R3

I2

6V1 + V2 = 32. Figure 3-7: Circuit containing two meshes with mesh currents Additionally, the supernode KVL equation is

I1 and I2 .

V2 − V1 = 18. Simultaneous solution of the two equations yields V1 = 2 V,

V2 = 20 V.

Concept Question 3-3: What impact does the presence

of a dependent source have on the implementation of the node-voltage method? (See )

Concept Question 3-4: What is a supernode? How is it

treated in nodal analysis? (See

)

Exercise 3-3: Apply the supernode concept to determine

I in the circuit of Fig. E3.3.

_

I

10 Ω



Ia = I1 . On the other hand, if an element is shared by two meshes, as is the case for R3 , the true branch current through it is the combination of the two branch currents: Ib = I1 − I2 .

12 V +

2A

mesh currents I1 and I2 . A mesh current may be thought of as the current flowing through the branches of that mesh, with no regard for the currents in neighboring meshes. That does not mean, however, that the mesh current is the same as the actual currents flowing through the elements of that mesh. For an element that belongs to only one mesh, such as R1 in Fig. 3-7, the current through it is indeed identical to the current in mesh 1. That is,

4Ω 4Ω

+ _ 20 V

Current I1 is assigned a positive sign because its direction through R3 is the same as that of Ib , but I2 is assigned a negative sign because it flows “upward” through R3 . The meshcurrent analysis method is based on the application of KVL to all of the meshes in the circuit. The solution procedure, which is analogous with that discussed earlier in Section 3-2 for the node-voltage method, consists of the following steps:

Figure E3.3 Answer: I = 0.5 A. (See

)

3-3 Mesh-Current Method 3-3.1

General Procedure

A mesh was defined in Section 1-3 as a loop that encloses no other loop. The current associated with a mesh is called its mesh current. The circuit in Fig. 3-7 contains two meshes with

Solution Procedure: Mesh Current Step 1: Identify all meshes and assign each of them an unknown mesh current. For convenience, define all mesh currents to be clockwise in direction. Step 2: Apply Kirchhoff’s voltage law (KVL) to each mesh. Step 3: Solve the resultant simultaneous equations to determine the mesh currents (see Appendix B).

“book” — 2015/5/4 — 7:08 — page 124 — #10

124

CHAPTER 3 ANALYSIS TECHNIQUES

For the circuit in Fig. 3-7, application of KVL to mesh 1, starting at the bottom left-hand corner and moving clockwise around the loop, gives −V0 + I1 R1 + (I1 − I2 )R3 = 0

(mesh 1),

(3.12)

where for each term we assigned a (+) or (−) sign to it depending on which of its voltage terminals is encountered first. Also, for a resistor, current flows into the (+) terminal of the voltage across it. For mesh 2, (I2 − I1 )R3 + I2 R2 = 0

(mesh 2).

R1

V0

I2 R3

=

V0

(mesh 1),

Sum of resistances Resistance shared in mesh 1 by meshes 1 and 2 + sign − sign Voltage source in mesh 1

(3.14a)

R3 I1

+

(R2 + R3) I2 = 0

I4

R4

R5

I3

R3

R6

Figure 3-8: Circuit for Example 3-4.

Solution: (a) Applying the symmetry pattern inherent in the structure of the mesh-current equations, we have (R1 + R2 + R5 )I1 − R2 I2 − R5 I3 = V0

and



I1

I2

(3.13)

The two simultaneous equations can be rearranged by collecting coefficients of I1 and I2 as

(R1 + R3) I1 −

+ _

R2

(mesh 2).

Resistance shared Sum of resistances by meshes 1 and 2 in mesh 2 − sign + sign

(3.14b) Note the built-in symmetry reflected by the structure of Eqs. (3.14a and b). For mesh 1, the coefficient of I1 in Eq. (3.14a) is the sum of all of the resistors contained in mesh 1, and the coefficient of I2 contains the resistor that mesh 1 shares with mesh 2. Furthermore, the coefficients of I1 and I2 have opposite signs. The same pattern applies for mesh 2 in Eq. (3.14b); the coefficient of I2 contains all of the resistors of mesh 2, and the coefficient of I1 contains the resistor shared by the two meshes. The magnitude of the voltage source in mesh 1 (namely, V0 ) appears on the right-hand side of Eq. (3.14a), with its polarity defined as positive if I1 flows through it from its negative to positive terminals. This structural pattern allows us to write the mesh-current equations directly, as discussed in more detail later in Section 3-4.

−R2 I1 + (R2 + R3 + R4 )I2 − R4 I3 = 0

(mesh 1), (3.15a) (mesh 2), (3.15b)

and −R5 I1 − R4 I2 + (R4 + R5 + R6 )I3 = 0

(mesh 3). (3.15c)

We note that in Eq. (3.15a) the coefficient of I1 is positive and is composed of the sum of all resistors in mesh 1 and the coefficients of I2 and I3 are negative and include the resistors that meshes 2 and 3 share with mesh 1, respectively. An equivalent pattern pertains to Eqs. (3.15b and c). If the mesh contains a voltage source, its magnitude appears on the right-hand side of the mesh equation and it is assigned a positive sign if it is a voltage rise when moving clockwise around the mesh. It is assigned a negative sign if it is a voltage drop. In the case of mesh 1 in the circuit of Fig. 3-8, V0 is a voltage rise, so it appears on the right-hand side of Eq. (3.15a) with a positive sign. (b) For the specified values of V0 and the six resistors, the three parts of Eq. (3.15) become 12I1 − 2I2 − 4I3 = 18,

Example 3-4: Circuit with Three Meshes

−2I1 + 8I2 − 4I3 = 0, −4I1 − 4I2 + 12I3 = 0,

Use mesh analysis to (a) obtain mesh-current equations for the circuit in Fig. 3-8 and then (b) determine the current in R4 , given that V0 = 18 V, R1 = 6 , R2 = R3 = 2 , and R4 = R5 = R6 = 4 .

and solution of the simultaneous equations leads to I1 = 2 A,

I2 = 1 A,

I3 = 1 A.

“book” — 2015/5/4 — 7:08 — page 125 — #11

3-3

MESH-CURRENT METHOD

125

The current through R4 is

Ix = 4V1

I4 = I3 − I2 = 1 − 1 = 0. Given that the circuit is a Wheatstone bridge (Section 2-5) operated under the balanced condition (R2 R6 = R3 R5 ), the result I4 = 0 is exactly what we should have expected. Exercise 3-4: Apply mesh analysis to determine I in the circuit of Fig. E3.4.

I3 1Ω

10 V

+ _

I1



+ _ 2Ω

V1

I2



4Ω I=? 12 V

+ _



3A

Figure 3-9: Mesh-current solution for a circuit containing a dependent source (Example 3-5).

Hence, I3 = 4V1 = 8(I1 − I2 ).

Figure E3.4 Answer: I = 0. (See

3-3.2

After inserting Eq. (3.17) into Eqs. (3.16a and b) and collecting terms in I1 and I2 , we end up with

)

−5I1 + 6I2 = 10,

Circuit with Dependent Sources

The presence of a dependent source in a circuit does not alter the basic procedure of the mesh-current method, but it requires the addition of a supplemental equation expressing the relationship between the dependent source and the other parts of the circuit.

Use mesh-current analysis to determine the magnitude of the dependent source Ix in Fig. 3-9. Solution: For the meshes with mesh currents I1 and I2 , (mesh 1),

−10I1 + 14I2 = 0. Solution of this pair of simultaneous equations gives I1 = −14 A,

Ix = 8(I1 − I2 ) = 8(−14 + 10) = −32 A. Exercise 3-5: Determine the current I in the circuit of

Fig. E3.5.

(3.16a)





and −2I1 + (2 + 1 + 3)I2 − I3 = 0

(mesh 2).

I2 = −10 A.

Hence,

Example 3-5: Dependent Current Source

(1 + 2)I1 − 2I2 − I3 = 10

(3.16b)

For mesh 3, we do not need to write a mesh-current equation, because I3 is specified by the current source as

60 V

I

I1

+ _

20 Ω

I3 = Ix = 4V1 . The voltage V1 across the 2  resistor is given by V1 = 2(I1 − I2 ).

(3.17)

Figure E3.5 Answer: I = 1.5 A. (See

)

I1 2

“book” — 2015/5/4 — 7:08 — page 126 — #12

126

TECHNOLOGY BRIEF 6: MEASUREMENT OF ELECTRICAL PROPERTIES OF SEA ICE

Technology Brief 6 Measurement of Electrical Properties of Sea Ice Climate change is often first measured by the decrease of our polar ice caps. This sea ice is a unique and vibrant type of ice; the fresh water freezes first, leaving pockets of more and more briny (salty) water, that eventually freezes only when the temperature gets below its eutectic point around −21 ◦ C. A combination of gravity and freeze-thaw cycles elongates these tiny brine pockets (initially sub-mm in size), and many of them start linking together to form fluidic channels (which eventually expand to become a full centimeter or more in diameter), from the top of the ice all the way through one or two meters of ice to the sea below the ice pack (Fig. TF6-1). In this columnar type of sea ice, which is prevalent in the Arctic, there is a critical brine volume fraction of about 5%, called the percolation threshold, above which there are largescale connected channels or pathways through which fluid can flow, and below which the sea ice is effectively impermeable. For a typical bulk sea-ice salinity of 5 parts per thousand, this brine volume fraction corresponds to a critical temperature of about −5 ◦ C. This on-off switch for fluid flow is known as the rule of fives.The brine channels can moderate the formation of melt ponds (Fig. TF6-2) by

quickly draining them and returning the ice to its more reflective white coloring. This brine percolation threshold has been quantified through measurements of the electrical resistivity of the ice, as well as X-ray computed tomography and measurements of the fluid permeability. Salty brine pockets are very conductive, and the surrounding ice is a near insulator. As the brine pockets join into channels, the overall conductivity of the ice increases substantially by providing a conducting path for current in pretty much the same way it provides a path for the water to percolate (drain) through. Conductivity, then, is highly correlated with the percolation threshold and can be used to help us study melt-pond formation. The electrical properties of the ice are measured by drilling out a 9 cm cylindrical core of ice, measuring its resistance using a model very similar to that seen in Fig. 2-1. Stainless steel nails are driven into the ice core (drilling holes for them first, to avoid cracking the core) to make the electrical connection to the ice. But this method has a problem. It is very hard to get a consistent electrical connection between the nail and the ice. This contact resistance is very much a part of the circuit, and it varies with each connection. A circuit model of this resistance measurement is shown below. The total resistance is the series combination of the two (variable) contact resistances and the resistance of the

Figure TF6-1: X-ray CT images (approximately 1 cm across) of the brine microstructure of sea ice. The brine volume fraction is 5.7%,

and the temperature is −8 ◦ C. Channels are beginning to form but are not fully connected yet. (From Golden et al., Geophys. Res. Letters, 2007.)

“book” — 2015/5/4 — 7:08 — page 127 — #13

TECHNOLOGY BRIEF 6: MEASUREMENT OF ELECTRICAL PROPERTIES OF SEA ICE

127

Ammeter Rcontact Voltmeter

Rsea ice

Rcontact Rsubject =

Voltmeter indication Ammeter indication

Figure TF6-4: 4-wire measurement circuit. Figure TF6-2: As ice melts, the liquid water collects in depressions on the surface and deepens them, forming these melt ponds in the Arctic. These fresh water ponds are separated from the salty sea below and around it, until breaks in the ice merge the two.

ice. Without being able to better control the contact resistance, Rsea ice cannot be accurately measured. To solve this problem, rather than doing a simple 2-wire resistance measurement as shown in Fig. TF6-3, a 4-wire measurement system can be used as shown in Fig. TF6-4. This system employs both an ammeter and a voltmeter (which are combined into the single yellow AEMC resistance meter shown in Fig. TF6-5). Two wires are used to connect the ammeter in series with the resistances, and two are used to connect the voltmeter in parallel with Rsea ice (hence, 4 wires). We do not need to know the driving voltage or the contact resistances in order to accurately measure Rsea ice with this method.

Rcontact Ohmmeter

Rsea ice

Rcontact Ohmmeter indicates Rcontact + Rsea ice + Rcontact FigureTF6-3: Simple 2-wire resistance measurement circuit.

FigureTF6-5: University of Utah mathematics Ph.D. student Christian Sampson measures the electrical conductivity of a seaice core during the Sea Ice Physics and Ecosystem eXperiment in 2012. Electrical clamps are attached to nails inserted along c Wendy Pyper/Australian Antarctic the length of the ice core. ( Division.)

“book” — 2015/5/4 — 7:08 — page 128 — #14

128

CHAPTER 3 ANALYSIS TECHNIQUES

3-3.3

Supermeshes

mesh currents, namely

 Two adjoining meshes that share a current source constitute a supermesh. The current source may be of the independent or dependent type, and it may include a resistor in series with it, but not in parallel.  The presence of a supermesh in a circuit, such as the one shown in Fig. 3-10(a), simplifies the solution by (a) combining the two mesh-current equations into one and (b) adding a simpler, auxiliary equation that relates the current of the source to the mesh currents of the two meshes. In Fig. 3-10(b), the current source of the supermesh has been removed (as has the series resistor R4 ) and replaced with a dashed line. The dashed line is a reminder to relate I0 to the

R1

Supermesh I2

R2 V0

+ _

I0

R4

I1 R5

I3

R3

R6

(a) Two adjoining meshes sharing a current source constitute a supermesh.

I0 = I2 − I3

V0

+ _

I2

R3

(R1 + R2 + R5 )I1 − R2 I2 − R5 I3 = V0 (mesh 1),

R5

Supermesh (b) Meshes 2 and 3 can be combined into a single supermesh equation, plus an auxiliary equation I0 = I2 − I3. Figure 3-10: Concept of a supermesh.

R6

(3.19)

and −(R2 + R5 )I1 + (R2 + R3 )I2 + (R5 + R6 )I3 = 0 (supermesh).

(3.20)

The two mesh-current equations, together with the auxiliary equation given by Eq. (3.18), are sufficient to solve for the three mesh currents. It is instructive to note that the series resistor R4 played no role in the solution. This is because the current through it is specified by I0 , regardless of the magnitude of R4 (so long as it is not an open circuit). Example 3-6: Circuit with a Supermesh

For the circuit in Fig. 3-11(a), determine (a) the mesh currents and (b) the power supplied by each of the two sources. Solution: (a) Meshes 3 and 4 share a current source, thereby forming a supermesh. Figure 3-11(b) shows the circuit redrawn such that meshes 3 and 4 can be combined into a single supermesh equation. Consequently, the mesh-current equations for mesh 1, mesh 2, and supermesh 3 and 4 respectively, are (10 + 2 + 4)I1 − 2I2 − 4I3 = 6

(mesh 1), (3.21a)

−2I1 + (2 + 2 + 2)I2 − 2I4 = 0

(mesh 2), (3.21b)

−4I1 − 2I2 + 4I3 + (2 + 4)I4 = 0

(supermesh). (3.21c)

and

I1 I3

(3.18)

The mesh-current equations for mesh 1 and the joint combination of meshes 2 and 3 are

R1 R2

(auxiliary eq.).

The auxiliary equation associated with the current source is given by I4 − I3 = 3

(auxiliary equation).

(3.22)

Inserting Eq. (3.22) to eliminate I4 in Eqs. (3.21b and c) leads to 16I1 − 2I2 − 4I3 = 6, −2I1 + 6I2 − 2I3 = 6, −4I1 − 2I2 + 10I3 = −18.

“book” — 2015/5/4 — 7:08 — page 129 — #15

3-4

BY-INSPECTION METHODS

129 (b) Since I1 = 0, the power supplied by the 6 V source is

10 Ω

P1 = 6I1 = 0.



I2

2Ω To calculate the power supplied by the 3 A current source, we need to know the voltage V1 across it, which is also the voltage across the 4  resistor given as

2Ω I1 6V

+ _ + V1 _



I4

I3

12 V1 = 4(I1 − I3 ) = 4 0 − − 7

3A





 =

48 V. 7

4Ω Hence, P2 = 3V1 = 3 ×

48 = 20.6 W. 7

Thus, all of the power is supplied by the 3 A source alone and is dissipated in the circuit resistances, except for the 10  resistance (because the current through it is I1 = 0).

(a) Original circuit 10 Ω

Concept Question 3-5: How does the presence of



a dependent source in the circuit influence the implementation procedure of the mesh-current method? (See )



I2 2Ω

I1 6V

Concept Question 3-6: What is a supermesh, and how is

+ _

I3 + V1 _



it used in mesh analysis? (See

I4 3A



)

Exercise 3-6: Apply mesh analysis to determine I in the circuit of Fig. E3.6.

I

3Ω Supermesh



4A

3A



(b) Meshes 3 and 4 constitute a supermesh Figure E3.6

Figure 3-11: Using the supermesh concept to simplify solution of the circuit in Example 3-6.

Solution of the three simultaneous equations gives 3 I2 = A, I1 = 0, 7 12 9 I3 = − A, I4 = A. 7 7

Answer: I = −0.7 A. (See

3-4

)

By-Inspection Methods

The node-voltage and mesh-current methods can be used to analyze any planar circuit, including those containing dependent sources. The solution process relies on the

“book” — 2015/5/4 — 7:08 — page 130 — #16

130

CHAPTER 3 ANALYSIS TECHNIQUES

application of KCL and KVL to generate the requisite number of equations necessary to solve for the unknown currents and voltages.

R2

Ia

 For circuits that contain only independent sources, their KCL and KVL equations exhibit standard patterns, allowing us to write them down by direct inspection of the circuit. The method of nodal analysis by inspection is easy to implement, but it requires that all sources in the circuit be independent current sources. Similarly, mesh analysis by inspection requires that all sources be independent voltage sources. 

V1

R1

R3

Ib

(a) Original circuit

If a circuit contains a mixture of independent current and voltage sources, implementation of the by-inspection methods will require a prerequisite step in which current sources are converted to voltage sources, or vice versa, so as to secure the requirement that all sources exclusively are current sources or voltage sources. The conversion process can be realized with the help of the source-transformation technique of Section 2-3.4.

3-4.1

V2

G2

Ia V1

V2

Nodal Analysis by Inspection G1

Even though it is common practice to characterize the i–υ relationship of a resistor in terms of its resistance R, it is more convenient in some cases to work in terms of its conductance G = 1/R and to apply the form of Ohm’s law given by I=

Ib

(b) Circuit in terms of conductances

V = GV . R

Figure 3-12: Application of the nodal-analysis by-inspection

The node-voltage by-inspection method is one such case. We shortly will demonstrate the method for the general case of a circuit composed of n (nonreference) extraordinary nodes. As noted earlier, applicability of the method is limited to circuits with independent current sources. By way of introducing the method, let us consider the simple circuit of Fig. 3-12(a), whose resistances have been relabeled in terms of conductances in Fig. 3-12(b). In a circuit diagram, the value next to the symbol of a resistor may be designated in ohms () or siemens (S), with the former referring to the value of its resistance R and the latter referring to the value of its conductance G. Both designations convey the same information about the resistor. The circuit has two extraordinary nodes. According to the node-voltage by-inspection method, the circuit is characterized by two node-voltage equations given by G11 V1 + G12 V2 = It1 ,

G3

(3.23a)

method is facilitated by replacing resistors with conductances.

and G21 V1 + G22 V2 = It2 ,

(3.23b)

where G11 and G22 = sum of all conductances connected to nodes 1 and 2, respectively G12 = G21 = negative of the sum of all conductances connected between nodes 1 and 2 It1 and It2 = total of all independent current sources entering nodes 1 and 2, respectively (a negative sign applies to a current source leaving a node).

“book” — 2015/5/4 — 7:08 — page 131 — #17

3-4

BY-INSPECTION METHODS

131

Application of these definitions to Fig. 3-12(b) gives

4A

G11 = G1 + G2 , G22 = G2 + G3 ,

V3

G12 = G21 = −G2 , It1 = −Ia ,

10 Ω (0.1 S)

and It2 = Ia + Ib .

V1

Hence, (G1 + G2 )V1 − G2 V2 = −Ia

V4

20 Ω (0.05 S)

10 Ω (0.1 S) V2

5Ω (0.2 S)

(3.24a)

2A

and −G2 V1 + (G2 + G3 )V2 = Ia + Ib .

1Ω (1 S)

2Ω (0.5 S)

3A

(3.24b)

It is a straightforward task to ascertain that Eqs. (3.24a and b) are indeed the correct node-voltage equations for the circuit in Fig. 3-12(b). Figure 3-13: Circuit for Example 3-7. Generalizing to the n-node case, the node-voltage equations can be cast in matrix form as ⎡

G11 ⎢ G21 ⎢ ⎢ .. ⎣ .

Gn1

G12 G22

··· ···

Gn2

···

⎤⎡ ⎤ ⎡ ⎤ G1n V1 It1 ⎢ ⎥ ⎢ ⎥ G2n ⎥ ⎥ ⎢ V2 ⎥ ⎢ It2 ⎥ ⎥ ⎢ .. ⎥ = ⎢ .. ⎥ , (3.25) ⎦⎣ . ⎦ ⎣ . ⎦

Gnn

Vn

Itn

Example 3-7: Four-Node Circuit

and abbreviated as GV = It ,

(3.26)

where G is the conductance matrix of the circuit, V is an unknown voltage vector representing the node voltages, and It is the source vector. The elements of these matrices are defined as Gkk = sum of all conductances connected to node k Gk = Gk = negative of conductance(s) connecting nodes k and , with k  =  (Gk = 0 if no conductance connects nodes k and  directly) Vk = voltage at node k Itk = total of current sources entering node k (a negative sign applies to a current source leaving the node). Solution of Eq. (3.26) for the elements of vector V can be obtained through matrix inversion (Appendix B) or the application of MATLAB or MathScript (Appendix E).

Obtain the node-voltage matrix equation for the circuit in Fig. 3-13 by inspection. Solution: At node 1, G11 =

1 1 1 + + = 1.3 S. 1 5 10

Similarly, at nodes 2, 3, and 4, 1 1 1 + + = 0.8 S, 5 2 10 1 1 = + = 0.15 S, 10 20

G22 = G33 and G44 =

1 1 + = 0.15 S. 10 20

“book” — 2015/5/4 — 7:08 — page 132 — #18

132

CHAPTER 3 ANALYSIS TECHNIQUES

The off-diagonal elements of the matrix are

1 = −0.2 S, 5 1 =− = −0.1 S, 10 = 0,

Exercise 3-7: Apply the node-analysis by-inspection

method to generate the node-voltage matrix for the circuit in Fig. E3.7.

G12 = G21 = − G13 = G31 G14 = G41

V1 2Ω

4A

G23 = G32 = 0, G24 = G42

1 = −0.1 S, =− 10

3A



Figure E3.7 Answer:



and G34 = G43

V2



1 =− = −0.05 S. 20

(See

5 6

− 13

− 13

8 15



V1 4 = . −3 V2

)

The total currents entering nodes 1 to 4 are

3-4.2

Mesh Analysis by Inspection

It1 = 2 A, It2 = 3 A, It3 = 4 A, and

By analogy with the node-voltage by-inspection method, for a circuit containing only independent voltage sources, its n mesh-current equations can be cast in matrix form as

It4 = −4 A.

Hence, the node-voltage matrix equation is given by ⎡

1.3 ⎢ −0.2 ⎢ ⎣ −0.1 0

−0.2 0.8 0 −0.1

−0.1 0 0.15 −0.05

⎤ ⎤⎡ ⎤ ⎡ 2 0 V1 ⎥ ⎢ ⎥ ⎢ −0.1 ⎥ ⎥ ⎢ V2 ⎥ = ⎢ 3 ⎥ , ⎦ ⎣ ⎦ ⎣ 4⎦ V3 −0.05 −4 V4 0.15

RI = Vt ,

(3.27)

where R is the resistance matrix of the circuit, I is a vector representing the unknown mesh currents, and V is the source vector. Equation (3.27) is an abbreviation for ⎡

R11 ⎢ R21 ⎢ ⎢ .. ⎣ .

Rn1

R12 R22

··· ···

Rn2

···

⎤⎡ ⎤ ⎡ ⎤ R1n I1 Vt1 ⎢ I2 ⎥ ⎢ Vt2 ⎥ R2n ⎥ ⎥⎢ ⎥ ⎢ ⎥ ⎥ ⎢ .. ⎥ = ⎢ .. ⎥ , (3.28) ⎦⎣ . ⎦ ⎣ . ⎦

Rnn

In

Vtn

where Solution by matrix inversion or MATLAB or MathScript software gives

V1 = 3.73 V, V2 = 2.54 V, V3 = 23.43 V, V4 = −17.16 V.

Rkk = sum of all resistances in mesh k, Rk = Rk = negative of the sum of all resistances shared between meshes k and  (with k = ) (Rk = 0 if meshes k and  do not share a resistor). Ik = current of mesh k Vtk = total of all independent voltage sources in mesh k, with positive assigned to a voltage rise when moving around the mesh in a clockwise direction.

“book” — 2015/5/4 — 7:08 — page 133 — #19

3-5

LINEAR CIRCUITS AND SOURCE SUPERPOSITION



I2





133

Concept Question 3-8: If the circuit contains a mixture of real voltage and current sources, what step should be taken to prepare the circuit for application of one of the two by-inspection methods? (See )



6V

+ _

I1

Exercise 3-8: Use the by-inspection method to generate the mesh-current matrix for the circuit in Fig. E3.8.



4V

+_ I3

+ _

12 Ω



2V

5Ω 10 Ω

+ 4V _

I1

_ +

8V

I3

20 Ω I2



Figure 3-14: Three-mesh circuit of Example 3-8. Figure E3.8 Example 3-8: Three-Mesh Circuit Answer:



Obtain the mesh-current matrix equation for the circuit in Fig. 3-14, by inspection. Solution: Application of the definitions for the elements of the matrix R and vector Vt leads to ⎡ ⎤⎡ ⎤ (2 + 3 + 6) −3 −6 I1 ⎣ ⎦ ⎣ I2 ⎦ −3 (3 + 4 + 5) −5 I3 −6 −5 (5 + 6 + 7) ⎡ ⎤ 6−4 = ⎣ 0 ⎦, 4 which simplifies to ⎡ 11 −3 ⎣ −3 12 −6 −5

⎡ ⎤ 2 −6 I1 −5 ⎦ ⎣ I2 ⎦ = ⎣ 0 ⎦ . 4 I3 18 ⎤⎡



Solution of the matrix equation gives I1 = 0.55 A, I2 = 0.35 A, and I3 = 0.50 A. Concept Question 3-7: Are the by-inspection methods

applicable to (a) circuits containing a mixture of independent voltage and current sources or (b) circuits containing a mixture of independent and dependent voltage sources? (See )

15 ⎣ −10 0

(See

3-5

−10 36 −20

⎤⎡ ⎤ ⎡ ⎤ 0 I1 12 −20 ⎦ ⎣ I2 ⎦ = ⎣ −8 ⎦ 32 −2 I3

)

Linear Circuits and Source Superposition

 A system is said to be linear if its output response is directly proportional to the excitation at its input.  In the case of a resistive circuit, the input excitation consists of the combination of all independent voltage and current sources in the circuit, and the output response consists of the set of all voltages across all passive elements in the circuit (namely, the resistors), or all currents through them. As noted in Section 3-1, circuits with ideal elements (including those containing capacitors and inductors) satisfy the linearity property, and therefore qualify as linear systems. A linear system obeys the superposition principle (Section 3-1.2), which for a linear circuit translates into:

“book” — 2015/5/4 — 7:08 — page 134 — #20

134

CHAPTER 3 ANALYSIS TECHNIQUES

 If a circuit contains more than one independent source, the voltage (or current) response of any element in the circuit is equal to the algebraic sum (superposition) of the individual responses associated with the individual independent sources, as if each had been acting alone.  Thus, for a circuit with n independent voltage or current sources labeled as sources 1 to n, the voltage υ across a given passive circuit element is given by υ = υ 1 + υ 2 + · · · + υn ,

(3.29)

Example 3-9: Circuit Analysis by Source Superposition

(a) Use source superposition to determine the current I in the circuit of Fig. 3-15. (b) Determine the amount of power dissipated in the 10  resistor due to each source acting alone and due to both sources acting simultaneously. Solution: (a) The circuit contains two sources, I0 and V0 . We start by transforming the circuit into the sum of two new circuits (one with I0 alone and another with V0 alone), as shown in parts (b) and (c) of Fig. 3-15, respectively. The current through R2 due to I0 alone is labeled I1 , and that due to V0 alone is labeled I2 .

where υk is the response when all sources have been set to zero, except for source k. A similar expression applies to the current i through the circuit, i = i1 + i2 + · · · + in .

I

R2 = 10 Ω

(3.30)

I0 = 6 A

The superposition principle can be used to find υ (or i) by executing the following steps:

Solution Procedure: Source Superposition

I1

Step 2: Apply node-voltage, mesh-current, or any other convenient analysis technique to solve for the response υ1 due to source 1 acting alone.

I0 = 6 A

(b)

Step 4: Use Eq. (3.29) to determine the total response υ. Alternatively, the procedure can be used to find currents i1 to in and then to add them up algebraically to find the total current i using Eq. (3.30). Because it entails solving a circuit multiple times, the sourcesuperposition method may not seem attractive, particularly for analyzing circuits with many sources. However, it is a useful tool in both analysis and design for evaluating the sensitivity of a response (such as the current in a load resistor) to specific sources in the circuit.  Whereas the source-superposition method is applicable for calculating voltage and current, it is not applicable for power (see Example 3-9). 

V0 = 45 V

+

+ -_

(a) Original circuit

Step 1: Set all independent sources equal to zero (by replacing voltage sources with short circuits and current sources with open circuits), except for source 1.

Step 3: Repeat the process for sources 2 through n, calculating in each case the response due to that one source acting alone.

R1 = 5 Ω

R2 = 10 Ω

R1 = 5 Ω

Source I0 alone generates I1 [Eliminating a voltage source = replacing it with short circuit] I2

R2 = 10 Ω

R1 = 5 Ω

(c)

V0 = 45 V

+

+ -_

Source V0 alone generates I2 [Eliminating a current source = replacing it with open circuit]

Figure 3-15: Application of the source-superposition method to the circuit of Example 3-9.

“book” — 2015/5/4 — 7:08 — page 135 — #21

3-5

LINEAR CIRCUITS AND SOURCE SUPERPOSITION

135

Circuit with current source alone Setting V0 = 0 means replacing the voltage source with a short circuit, as shown in Fig. 3-15(b). By current division,     R1 5 I0 = 6 = 2 A. I1 = R1 + R 2 5 + 10

2Ω 9Ω

+ 6V _

+ _

2V



2Vx

+

Vx



Circuit with current source alone Setting I0 = 0 means replacing the current source with an open circuit, as shown in Fig. 3-15(c). Application of KVL leads to   −45 V0 = = −3 A. I2 = − R1 + R 2 5 + 10 Hence, I = I1 + I2 = 2 − 3 = −1 A.

(a) Original circuit Vx1

2Ω 9Ω

+ 6V _



2Vx1

+

Vx1



(b) The amounts of power dissipated in the 10  resistor due to I1 alone, I2 alone, and the total current I are, respectively; P1 = I12 R = 22 × 10 = 40 W,

(b) The 6 V source acting alone generates voltage Vx1

P2 = I22 R = (−3)2 × 10 = 90 W,

Vx2



and



P = I 2 R = 12 × 10 = 10 W. Note that P  = P1 + P2 , because the linearity property does not apply to power. Example 3-10: Superposition for Dependent-Source Circuit

Apply the superposition principle to the circuit shown in Fig. 3-16(a) to determine Vx . Solution: The circuit in Fig. 3-16(a) contains two independent voltage sources. Our task is to determine voltage Vx across the 6  resistor. The superposition method allows us to represent the original circuit by two new circuits, one containing the 6 V source while excluding the 2 V source, and another with the opposite arrangement. The first circuit generates Vx1 across the 6  resistor and the second circuit generates Vx2 . The unknown voltage Vx is the sum of the two.

2V

+ _

2Vx2

Vx1 − 6 Vx1 Vx + − 2Vx1 + 1 = 0, 2 9 6 which leads to Vx1 = −2.45 V.



(c) The 2 V source acting alone generates voltage Vx2 Figure 3-16: Application of superposition to the circuit of Example 3-10.

Circuit with 2 V source alone At node Vx2 in the circuit of Fig. 3-16(c), KCL gives Vx2 Vx Vx − 2 + 2 − 2Vx2 + 2 = 0, 2 9 6

Circuit with 6 V source alone At node Vx1 in the circuit of Fig. 3-16(b), KCL gives

+

Vx2

which leads to Vx2 = −0.18 V. Hence, Vx = Vx1 + Vx2 = −2.45 − 0.18 = −2.63 V.

“book” — 2015/5/4 — 7:08 — page 136 — #22

136

TECHNOLOGY BRIEF 7: INTEGRATED CIRCUIT FABRICATION PROCESS

Technology Brief 7 Integrated Circuit Fabrication Process Do you ever wonder how the processor in your computer was actually fabricated? How is it that engineers can put hundreds of millions of transistors into one device that measures only a few centimeters on a side (and with so few errors) so the devices actually function as expected? Devices such as modern computer processors and semiconductor memories fall into a class known as integrated circuits (IC). They are so named because all of the components in the circuit (and their “wires”) are fabricated simultaneously (integrated) onto a circuit during the manufacturing process. This is in contrast to circuits where each component is fabricated separately and then soldered or wired together onto a common board (such as those you probably build in your lab classes). Integrated circuits were first demonstrated independently by Jack Kilby at Texas Instruments and Robert Noyce at Fairchild Semiconductor in the late 1950s. Once developed, the ability to easily manufacture components and their connections with good quality control meant that circuits with thousands (then millions, then billions) of components could be designed and built reliably.

Figure TF7-1: A single 4-inch silicon wafer. Note the wafer’s mirror-like surface. (Courtesy of Veljko Milanovic.)

Semiconductor Processing Basics All mainstream semiconductor integrated-circuit processes start with a thin slice of silicon, known as a substrate or wafer.This wafer is circular and ranges from 4 to 18 inches in diameter and is approximately 1 mm thick (hence its name). Each wafer is cut from a single crystal of the element silicon and polished to its final thickness with atomic smoothness (Fig.TF7-1). Most circuit designs (like your processor) fit into a few square centimeters of silicon area; each self-contained area is known as a die. After fabrication, the wafer is cut to produce independent, rectangular dies often known as chips, which are then packaged with plastic covers and metal pins or other external connections to produce the final component you buy at the store. A specific sequence or process of chemical and mechanical modifications is performed on certain areas of the wafer. Although complex processes employ a variety of techniques, a basic IC process will employ one of the following three modifications to the wafer:

• Implantation: Atoms or molecules are driven into (implanted in) the silicon wafer, changing its electronic properties (Fig. TF7-2(a)). • Deposition: Materials such as metals, insulators, or semiconductors are deposited in thin layers (like spray painting) onto the wafer (Fig. TF7-2(b)). • Etching: Material is removed from the wafer through chemical reactions or mechanical motion (Fig. TF7-2(c)).

Lithography When building an IC, we need to perform different modifications to different areas of the wafer. We may want to etch some areas and add metal to others, for example. The method by which we define which areas will be modified is known as lithography.

“book” — 2015/5/4 — 7:08 — page 137 — #23

TECHNOLOGY BRIEF 7: INTEGRATED CIRCUIT FABRICATION PROCESS

(a) Implantation: High-energy ions are driven into the silicon. Most become lodged in the first few nanometers, with decreasing concentration away from the surface. In this example, boron (an electron donor) is implanted into a silicon substrate to make a p-type material.

(b) Deposition: Atoms (or molecules) impact the surface but do not have the energy required to penetrate the surface. They accumulate on the surface in thin films. In this example, aluminum is deposited in a conductive film onto the silicon.

137

(a) Etching: Chemical, mechanical, or high-energy plasma methods are used to remove silicon (or other material) from the surface. In this example, silicon is etched away from the substrate.

Figure TF7-2: Cross-section of basic fabrication processes. The dashed line in each drawing indicates the original surface of the wafer.

Lithography has evolved much over the last 40 years and will continue to do so. Modern lithography employs all of the basic principles described below, but uses complex computation, specialized materials, and optical devices to achieve the very high resolutions required to reach modern feature sizes. At its heart, lithography is simply a stencil process. In an old-fashioned stencil process, when a plastic sheet with cut-out letters or numbers is laid on a flat surface and painted, the cutout areas will be painted and the rest will be masked. Once the stencil is removed, the design left behind consists of only the painted areas with clean edges and a uniform surface. The total surface area of the IC depends on the number and complexity of the circuit elements on the IC, and on the minimum feature size, which is 10 nm (10−8 m) today. With that in mind, consider Fig. TF7-3. Given a flat wafer, we first apply a thin coating of liquid polymer known as photoresist (PR). This layer usually is several hundred nanometers thick and is applied by placing a drop in the center of the wafer and then spinning the wafer very fast (1000 to

5000 rpm) so that the drop spreads out evenly over the surface. Once coated, the PR is heated (usually between 60 to 100 ◦ C) in a process known as baking; this allows the PR to solidify slightly to a plastic-like consistency. This layer is then exposed to ultraviolet (UV) light, the bonds that hold the PR molecules together are “chopped” up; this makes it easy to wash away the UV-exposed areas (some varieties of PR behave in exactly the opposite manner: UV light makes the PR very strong or cross-linked, but we will ignore that technique here). In lithography, UV light is focused through a glass plate with patterns on it; this is known as exposure. These patterns act as a “light stencil” for the PR.Wherever UV light hits the PR, that area subsequently can be washed away in a process called development. After development, the PR film remains behind with holes in the exposed and washed areas. How is this helpful? Let’s look at how the modifications presented earlier can be masked with PR to produce patterned effects (Fig. TF7-4). In each case, we first use lithography to pattern areas onto the wafer (Fig. TF7-4(a)) then we perform one of our three

“book” — 2015/5/4 — 7:08 — page 138 — #24

138

TECHNOLOGY BRIEF 7: INTEGRATED CIRCUIT FABRICATION PROCESS

Dispense

Spin and Bake

Expose

Develop

Figure TF7-3: Basic lithography steps.

processes (Fig. TF7-4(b)), and finally, we use a strong solvent such as acetone (nail polish remover) to completely wash away the PR (Fig. TF7-4(c)). The PR allows us to implant, deposit, or etch only in defined areas.

Fabricating a Diode In Section 2-6, we discussed the functional performance of the diode as a circuit component. Here, we will examine briefly how a diode is fabricated. Similar but

more complex multi-step processes are used to make transistors and integrated circuits. Conceptually, the simplest diode is made from two slabs of silicon— each implanted with different atoms—pressed together such that they share a boundary (Fig. TF7-5). The n and p areas are pieces of silicon that have been implanted with atoms (known as impurities) that change the number of electrons capable of flowing freely through the silicon. This changes the semiconducting properties of the silicon and creates an electrically active boundary

Lithography

Implantation

p or n type silicon area

Deposition

Metal film

Etch

Etched recess

silicon substrate

Figure TF7-4: Lithography used to pattern implantation areas, deposit metal features, and etch areas.

“book” — 2015/5/4 — 7:08 — page 139 — #25

TECHNOLOGY BRIEF 7: INTEGRATED CIRCUIT FABRICATION PROCESS

p-type silicon

139

n-type silicon

metal

metal

Figure TF7-5: The basic diode (top) circuit symbol and (bottom) conceptual depiction of the physical structure.

n-type implant

Lithography + etch oxide

a f Grow oxide Remove PR

b g Lithography + etch oxide

Metal deposition

c h

Remove PR

Lithography + etch metal

d i p-type implant

e Complete diode

j

Metal

Metal

Figure TF7-6: A simple pn-junction diode fabrication process.

Figure TF7-7: Colorized scanning electron-microscope cross section of a 64-bit high-performance microprocessor chip built in IBM’s 90 nm Server-Class CMOS technology. Note that several metal interconnect levels are used (metal lines are orange, insulator is green); the transistors lie below this metal on the silicon wafer itself (dark blue). (Courtesy of International Business Machines Corporation.)

(called a junction) between the n and the p areas of silicon. If a forward-biased voltage is applied, it is as if the p charges move towards the n side, allowing current to flow, even though no actual p or n atoms move in the diode. When both the n and p pieces of silicon are connected to metal wires, this two-terminal device exhibits the diode i–υ curve shown in Fig. 2-40(c). Figure TF7-6 shows the process for making a single diode. Only one step needs further definition: oxidation. During oxidation, the silicon wafer is heated to > 1000 ◦ C in an oxygen atmosphere. At this temperature, the oxygen atoms and the silicon react and form a layer of SiO2 on the surface (this layer is often called an oxide layer). SiO2 is a type of glass and is used as an insulator. Wires are made by depositing metal layers on top of the device; these are called interconnects. Modern ICs have 6 to 7 such interconnect layers (Fig. TF7-7). These layers are used to make electrical connections between all of the various components in the IC in the same way that macroscopic wires are used to link components on a breadboard.

“book” — 2015/5/4 — 7:08 — page 140 — #26

140

CHAPTER 3 ANALYSIS TECHNIQUES

Concept Question 3-9: Explain why the linearity

property of electric circuits is an underlying requirement for the application of the source-superposition method. (See ) Concept Question 3-10: How is the superposition

method used as a sensitivity tool in circuit analysis and design? (See ) Concept Question 3-11: Is the source-superposition

method applicable to power? In other words, if source 1 alone supplies power P1 to a certain device and source 2 alone supplies power P2 to the same device, will the two sources acting simultaneously supply power P1 + P2 to the device? (See ) Exercise 3-9: Apply the source-superposition method to determine the current I in the circuit of Fig. E3.9.

3Ω 2Ω

I

4A

3A



Figure E3.9 Answer: I = 2.3 A. (See

)

Exercise 3-10: Apply source superposition to deter-

mine Vout in the circuit of Fig. E3.10.



3A



4A



+

Vout



Figure E3.10 Answer: Vout = −1 V. (See

)

3-6 Th´evenin and Norton Equivalent Circuits As depicted by the block diagram shown in Fig. 3-17, a generic cell-phone circuit consists of several individual

circuits, including amplifiers, oscillators, analog-to-digital (A/D) and digital-to-analog (D/A) converters, an antenna, a diplexer that allows the antenna to be used for both transmission and reception, a microprocessor, and other auxiliary circuits. Many of these circuits are quite complex and may contain a large number of active and passive elements, in both discrete and integrated form. So the question one might ask is: How does an engineer approach an analysis or design task involving such a complex architecture? Dealing with the entire circuit all at once would be next to impossible, not only because of its daunting complexity, but also because the individual circuits call for engineers with different specializations. Fortunately, we have a straightforward answer to the question, namely that each circuit gets modeled as a “black box,” or block, with specified input and output (I/O) terminal characteristics allowing the engineer working with a particular circuit to treat the other circuits connected to it in terms of only those (I/O) characteristics without much regard to the details of their internal architectures. For an amplifier, for example, its overall specifications might include voltage gain and frequency bandwidth, among other attributes, but its terminal characteristics refer to how it would “appear” from the perspective of other circuits. Conversely, from the amplifier’s perspective, other circuits are specified in terms of how they appear to the amplifier. Figure 3-18 illustrates the concept from the perspective of the radio-frequency (RF) low-noise amplifier in the receive channel of the cell-phone circuit. The combination of the antenna and diplexer (including the input signal picked up by the antenna) is represented at the input side of the amplifier by an equivalent circuit composed of a voltage source υs in series with an impedance Zs . Impedance (which we shall introduce in a later chapter) is the ac-equivalent of resistance in dc circuits. At the output side of the amplifier, the mixer (whose function is to shift the center frequency of the input signal from 834 MHz down to 70 MHz) is represented by a load impedance ZL . Thus, the output terminal characteristics of the antenna/diplexer combination become the input source to the amplifier, and the input impedance of the mixer becomes the load to which the amplifier is connected.

 Isolating the amplifier, while keeping it in the context of its input and output neighbors, facilitates both the analysis and design processes. 

“book” — 2015/5/4 — 7:08 — page 141 — #27

´ 3-6 THEVENIN AND NORTON EQUIVALENT CIRCUITS

RF = Radio Frequency IF = Intermediate Frequency LO = Local Oscillator RF Power Mixer = Frequency Up- or Amp Down-Converter RF Filter Transmit Path

141

Human Interface, Dialing, Memory Battery Power Control Mixer

(Speech, video, data) In Out

Microprocessor Control

IF Amp Modulator

Antenna Transmitted Signal

LO

~

~ IF Amp

LO

D/A and A/D Converters and Filters

Demodulator Received Signal

Diplexer/Filter

RF Low Mixer Noise Amp

IF Filter

Receive Path Antenna and Propagation

RF Front-End

IF Block

Back-End

Baseband

Figure 3-17: Cell-phone block diagram.

3-6.1

Input and Output Resistances

Example 1: Household wiring Our homes are powered by some kind of electrical generation plant (coal-powered, hydroelectric, etc.) that produces high voltage, which is run to our city on high-voltage transmission lines, split into smaller voltages at various substations, and eventually delivered to the breakers or fuse boxes of our homes

(Chapter 10). This is a rather complex system with many parts, so we prefer not to analyze the entire system every time we consider a change in a household electrical circuit. We can represent the entire power distribution system as a voltage source (in this case, 110 V) in series with a small source resistance Rs that represents the losses in the transmission lines and connections, as shown in the simplified block diagram in Fig. 3-19. Even though the source is ac, we will treat it as if it were a dc source with Vs = 110 V.

Source impedance Zs υs

+ _

+

Zin

RF low-noise amplifier

υin

Zout

_ Input equivalent circuit

+ υout

ZL

_ Amplifier circuit

Mixer input impedance

Load equivalent circuit

Figure 3-18: Input and output circuits as seen from the perspective of a radio-frequency amplifier circuit.

“book” — 2015/5/4 — 7:08 — page 142 — #28

142

CHAPTER 3 ANALYSIS TECHNIQUES

Lamp

Breaker box

Fan

Power plant (a) Electrical system

I + Vs _ 110 V

Rs DVM Rfan

Rlamp

(b) Equivalent circuit

Figure 3-19: (a) Power distribution system driving a fan and a lamp in a house, and (b) block diagram of the source (power distribution system), fan, lamp, and a voltmeter measuring the voltage in the outlet.

Every device we plug in (such as the fan and lamp) is in parallel with the power source block. The lamp is just a switch and a light bulb, which we might even simplify further by ignoring the switch and assuming it is always on, thus giving us very simply a resistor Rlamp in the block diagram in Fig. 3-19(b). The fan, on the other hand, is a little more complicated because it includes a motor and a switch that controls various speeds, but we can still represent it by a parallel resistor Rfan . If Vs = 110 V, Rs = 10 , and Rfan = Rlamp = 100 , what is the current drawn from the source? The parallel combination of Rfan and Rlamp is R = Rfan  Rlamp = 100  100 = 50 . The total resistance connected to Vs is the series sum of R and Rs : Rtotal = R + Rs = 50 + 10 = 60 . Hence, by Ohm’s law, I=

110 Vs = 1.83 A. = Rtotal 60

What is the voltage across the fan and lamp? Application of voltage division gives

Vfan = Vlamp = Vs

R 110 × 50 = = 91.67 V. Rtotal 60

This is measurably less than the 110 V of Vs . The voltage reduction is called loading the circuit, and it occurs when the series source input resistance and the load resistance (Rfan  Rlamp ) are on the same order of magnitude, or if the source resistance is larger than the load resistance. If too many appliances are plugged into the outlet, all of their resistances combine in parallel, thereby reducing the total load resistance, drawing more current, and loading down the circuit (reducing the voltage across the devices). Eventually, the devices will no longer function properly (if the voltage gets too low) or the circuit breaker creates an open circuit if the current gets too high. This example illustrates the concept of input and output resistances and why they matter. The input resistance is what is seen looking into a block “from the left,” and the output resistance is what is seen looking in “from the right.” For the

“book” — 2015/5/4 — 7:08 — page 143 — #29

´ 3-6 THEVENIN AND NORTON EQUIVALENT CIRCUITS voltage source, we do not really have an input resistance, and its output resistance is Rs . For the fan and lamp, the input and output resistances are both Rfan and Rlamp , respectively. If we have a small output resistance looking into (connected to) a large input resistance of a load (such as the fan/lamp circuit), the load will not draw down the voltage (load the circuit) very much. In fact, if the input resistance of the load is high enough, we can even ignore it in the analysis of the circuit because the voltage across it is essentially the same as the source voltage. On the other hand, if the output and input resistances are similar in magnitude, the load will indeed draw down the voltage (and load the circuit). The load clearly has an impact on the circuit, and we cannot ignore it in the analysis of how the circuit works. And if the output resistance of the source is larger than the input resistance of the load, the voltage of the load will be significantly reduced (loaded).

Example 2: Voltmeter Voltmeters are deliberately designed with high input resistance (≥ 2 M) so that they do not affect the circuit being measured. Consider, for example, measuring the voltage (or resistance) across the fan/lamp circuit in Fig. 3-19(b). The resistance of the fan and lamp in parallel is 50 . If the voltmeter (DVM) has an input resistance of 2 M, the fan/lamp/DVM circuit has a total resistance of 49.999 , a change of less than 0.01%. Another way to think of this is that the DVM will draw very little current through it, because of its high input resistance. If the DVM is used to measure resistance instead of voltage, its input resistance also is high, and would have minimal effect on the circuit being measured. In contrast, the input resistance of an ammeter is very small (about 1 μ), much smaller than the fan/lamp combination. Summary: What have we learned from these examples? • Input resistance (looking toward the load) and output resistance (looking toward the source) are important parameters of the circuit. • If the input resistance of the load is very high compared with the output resistance of the rest of the circuit (such as the case with the voltmeter), that part of the circuit (the load) can basically be ignored when we analyze the other parts of the circuit. In fact, this means that these blocks can be designed and analyzed individually. We call them independent, uncoupled, or decoupled. Being able

143 to design and analyze blocks of a circuit individually is such a powerful concept that we often deliberately build circuits to have high input resistance. Circuits with high input resistance draw minimal current. • If the input resistance of the load is low (or similar) compared with the output resistance of the input circuit, significant current is drawn into the load circuit. This may load the source circuit and reduce the voltage at the load. Also, the circuits can no longer be analyzed individually; they are coupled and must be analyzed together.

3-6.2 Th´evenin’s Theorem  Our ability to develop equivalent-circuit representations is made possible (in part) by a pair of theorems of fundamental significance known as Th´evenin’s and Norton’s theorems. 

Most electrical systems are quite complex, so that each subsystem (such as the filter, demodulator, amplifier, etc., in Fig. 3-17) is designed independently, and often by different engineers and even by different companies. We established in the preceding subsection that in order to design subsystems independently, it is necessary that each has a high input resistance. This is often not feasible, however, so we need another approach to designing cascaded circuits and then combining them together into a larger system. The Th´evenin and Norton concepts described in this section help us do that. They are very powerful techniques used extensively in electrical engineering design. In practice, the system engineering team determines what blocks are needed for the system, lays out the block diagram, and specifies the input voltages and currents, and input and output resistances for each block of the circuit. Design teams then create circuits for each block, and test them independently using the input and output resistances/ voltages/currents for their neighboring subsystems in the test protocol. The integration team puts the subsystems together, often with the mechanical parts of the system as well, and then tests the overall system as an integrated unit to insure that its performance meets the design specifications. In the 1880s, a French engineer named L´eon Th´evenin introduced the concept known today as Th´evenin’s theorem, which asserts:

“book” — 2015/5/4 — 7:08 — page 144 — #30

144

CHAPTER 3 ANALYSIS TECHNIQUES

A linear circuit can be represented at its output terminals by an equivalent circuit consisting of a series combination of a voltage source υTh and a resistor RTh , where υTh is the open-circuit voltage at those terminals (no load) and RTh is the equivalent resistance between the same terminals when all independent sources in the circuit have been deactivated. RTh is the output resistance of the Th´evenin circuit. 

a

+

Actual circuit

υoc



3-6.3

Finding υTh

Th´evenin equivalency means that from the standpoint of the load RL , the two circuits in Fig. 3-20 are indistinguishable. For

a Actual circuit

iL RL

b

Load (a) Original circuit RTh υTh

a

iL

+ _ b

(b)

υTh

' Thevenin equivalent

Figure 3-20: A circuit can be represented in terms of a Th´evenin equivalent comprising a voltage source υTh in series with a resistance RTh .

a

+ _

+

υTh



+

_

b (b) Measuring υTh of equivalent circuit Figure 3-21: Equivalency means that υTh of the Th´evenin equivalent circuit is equal to the open-circuit voltage for the actual circuit.

any value we assign to RL , both circuits generate the same iL . Hence, if we disconnect RL altogether from both circuits and then measure the voltage across terminals (a, b), we should measure the same voltage. The scenario is depicted in Fig. 3-21. In part (a), a voltmeter would measure the open-circuit voltage υoc of the actual circuit, and in part (b) the voltmeter would measure υTh (since there is no voltage drop across RTh ). We are effectively measuring the output voltage of our blackbox. Equivalency requires that υTh (of Th´evenin equivalent) = υoc (of actual circuit). (3.31) The procedure is equally valid for circuits with or without dependent sources. For a circuit with no independent sources, υTh = 0.

3-6.4 RL

_

b (a) Measuring υoc on actual circuit RTh

A pictorial representation of Th´evenin’s theorem is shown in Fig. 3-20, where the actual circuit in part (a) has been replaced with the Th´evenin equivalent circuit in part (b). The implication of this model is that when the circuit is connected to a load resistor RL , the current iL running through it will be identical for both the actual circuit and the equivalent circuit. This equivalence holds true for any value of RL , from zero (short circuit) to ∞ (open circuit). Thus, from the standpoint of the load, the two circuits are indistinguishable. Even though the present discussion pertains to dc currents, the Th´evenin concept extends to ac circuits as well. We will revisit the concept in a future chapter for circuits containing capacitors and inductors.

+

Finding RTh —Short-Circuit Method

Multiple methods are available for finding the Th´evenin resistance RTh . We start with the short-circuit method. From Fig. 3-20(b), υTh iL = . (3.32) RTh + RL If RL = 0 (short-circuit load), we call iL the short-circuit current isc , which would be given by isc =

υTh . RTh

(3.33)

“book” — 2015/5/4 — 7:08 — page 145 — #31

´ 3-6 THEVENIN AND NORTON EQUIVALENT CIRCUITS

Open-Circuit / Short-Circuit Method a Actual circuit

+

υoc



+

Volts

_

145

Example 3-11: Open Circuit / Short Circuit Method

The input circuit to the left of terminals (a, b) in Fig. 3-23(a) is connected to a variable load resistor RL . Determine (a) the Th´evenin equivalent of the circuit to the left of terminals (a, b) and (b) use it to find the value of RL that will cause the magnitude of the current through it to be 0.5 A.

b (a) υTh = υoc 6Ω

isc Actual circuit

a



IL

Amps

24 V

+ _

12 Ω

RL

7A b

(b) RTh = υoc /isc

(a) Original circuit

Figure 3-22: Th´evenin voltage is equal to the open-circuit voltage and Th´evenin resistance is equal to the ratio of υoc to isc , where isc is the short-circuit current between the output terminals.

Vc



24 V

a



+

+ _

12 Ω

Voc = VTh

7A

_ b (b) Replacing RL with open circuit By analyzing the circuit configuration in Fig. 3-22(b) to find isc or, measuring isc with an ammeter, we can apply Eq. (3.33) to find RTh , RTh

υTh = . isc

(3.34)

The only potential problem with this type of measurement is that when short-circuiting the source circuit, the current threshold of the ammeter may be exceeded (if the output resistance of the source circuit is very small). This method is applicable to any circuit with at least one independent source, regardless of whether or not it contains dependent sources.

Vc



24 V

a



Isc

+ _

12 Ω

7A b

(c) Replacing RL with short circuit RTh = 6 Ω

a IL

VTh = 12 V

_ +

RL b

' (d) Thevenin equivalent circuit

Figure 3-23: Applying open circuit/short circuit method to find the Th´evenin equivalent for the circuit of Example 3-10.

“book” — 2015/5/4 — 7:08 — page 146 — #32

146

CHAPTER 3 ANALYSIS TECHNIQUES

Solution: (a) With RL replaced with an open circuit in Fig. 3-23(b), VTh is the open-circuit voltage between terminals (a, b). Since no current flows through the 2  resistor, VTh = Vc at node c. The node-voltage equation at node c is Vc − 24 Vc + + 7 = 0, 6 12

Exercise 3-11: Determine the Th´evenin-equivalent circuit at terminals (a, b) in Fig. E3.11.

3Ω 2Ω

4A

Next, we replace RL with a short circuit (Fig. 3-23(c)) and repeat the process to find Vc : V Vc − 24 Vc + + 7 + c = 0, 6 12 2 whose solution gives Vc = −4 V, and Vc 2 4 =− 2 = −2 A.

Isc =

Hence, VTh Isc −12 = −2

RTh =

= 6 , and the Th´evenin equivalent circuit is shown in Fig. 3-23(d). (b) In view of Fig. 3-23(d), for IL to be 0.5 A, it is necessary that IL =

12 6 + RL

3A

5Ω b

which leads to Vc = −12 V. Hence, VTh = −12 V.

a

Figure E3.11 Answer: VTh = −3.5 V, Isc = −1.4 A, RTh = 2.5 .

(See

3-6.5

)

Finding RTh —Equivalent Resistance Method

If the circuit does not contain dependent sources, RTh can be determined by deactivating all sources (replacing voltage sources with short circuits and current sources with open circuits) and then simplifying the circuit down to a single equivalent resistance between its output terminals, as portrayed by Fig. 3-24. In that case, RTh = Req .

(3.35)

This method does not apply to circuits that contain dependent sources. Example 3-12: Thevenin ´ Resistance

Find RTh at terminals (a, b) for the circuit in Fig. 3-25(a). Solution: Since the circuit has no dependent sources, we can apply the equivalent-resistance method. We start by

Equivalent-Resistance Method Circuit with all independent sources deactivated

Req = RTh

= 0.5 A Figure 3-24: For a circuit that does not contain dependent or RL = 18 .

sources, RTh can be determined by deactivating all sources (replacing voltage sources with short circuits and current sources with open circuits) and then simplifying the circuit down to a single resistance Req .

“book” — 2015/5/4 — 7:08 — page 147 — #33

´ 3-6 THEVENIN AND NORTON EQUIVALENT CIRCUITS

147 resultant 60  with the 30  resistance in parallel, we obtain

16 V _

RTh = 20 .

+

50 Ω (a) Original circuit 4A

50 Ω a 30 Ω

Exercise 3-12: Find the Th´evenin equivalent of the circuit

to the left of terminals (a, b) in Fig. E3.12, and then determine the current I .

b 35 Ω

5Ω 5Ω 20 V

50 Ω (b) After deactivating sources

+ _

I

b





5A

Figure E3.12 Answer:

30 Ω



b 2V

25 Ω 35 Ω

a

a

(See

a 30 Ω b

a (d) Final RTh



50 Ω

35 Ω

(c) After combining the two 50 Ω resistors in parallel

0.6 Ω

RTh = 20 Ω b

Figure 3-25: After deactivation of sources, systematic simplification leads to RTh (Example 3-12).

deactivating all of the sources (as shown in Fig. 3-25(b)) where we replaced the voltage source with a short circuit and the current source with an open circuit. After (a) combining the two 50  resistors in parallel, (b) combining their 25  combination in series with the 35  resistance, and (c) finally combining the

3-6.6

+ _

a

I

b

1Ω I = 0.5 A.

)

Finding RTh —External-Source Method

The equivalent-resistance method described previously does not apply to circuits containing dependent sources. Hence, an alternative variation is called for. Independent sources again are deactivated (but dependent sources are left alone) and an external voltage source υex is introduced to excite the circuit, as shown in Fig. 3-26. After analyzing the circuit to determine the current iex , RTh is found by applying RTh =

υex . iex

(3.36)

Since iex is caused by υex , it is directly proportional to it. Hence, we may choose any value for υex , such as υex = 1 V, as long as we use the same value both in Fig. 3-26 when analyzing the circuit to find iex and in applying Eq. (3.36) to compute RTh . Example 3-13: Circuit with Dependent Source

Find the Th´evenin equivalent circuit at terminals (a, b) for the circuit in Fig. 3-27(a) by applying the combination of opencircuit-voltage and external-source methods.

“book” — 2015/5/4 — 7:08 — page 148 — #34

148

CHAPTER 3 ANALYSIS TECHNIQUES



External-Source Method + _

Circuit with only independent sources deactivated

iex υex

a



+

Ix + 33 V _

2Ω Vab = VTh

I1

+ -_ 3Ix

− + _

RTh

(a) Solving for VTh

iex υex





b

a

Iex

Ix

Figure 3-26: If a circuit contains both dependent and



independent sources, RTh can be determined by (a) deactivating only independent sources ( by replacing independent voltage sources with short circuits and independent current sources with open circuits), (b) adding an external source υex , and then (c) solving the circuit to determine iex . The solution is RTh = υex /iex .

I1

+ -_ 3Ix

+ _

I2

(b) Solving for Iex Solution: The KVL equation for mesh current I1 in Fig. 3-27(a) is given by

b

a

−33 + 6I1 + 2I1 + 3Ix = 0.

Iex + _

RTh

Recognizing that Ix = I1 , solution of the preceding equation leads to I1 = 3 A.

Vex

Vex

b

Since there is no voltage drop across the 4  resistor (because no current is flowing through it),

(c) Equivalent circuit for calculating RTh

VTh = Vab = 2I1 + 3Ix = 5I1 = 15 V.

Figure 3-27: Solution of the open-circuit voltage gives

To find RTh using the external-source method, we deactivate the 33 V voltage source and we add an external voltage source Vex , as shown in Fig. 3-27(b). Our task is to obtain an expression for Iex in terms of Vex . In Fig. 3-27(b) we have two mesh currents, which we have labeled I1 and I2 . Their equations are given by 6I1 + 2(I1 − I2 ) + 3Ix = 0, −3Ix + 2(I2 − I1 ) + 4I2 + Vex = 0. After replacing Ix with I1 and solving the two simultaneous equations, we obtain I1 = −

1 Vex , 28

Vab = VTh = 15 V. Use of the external-voltage method leads to RTh = 56/11  (Example 3-13).

and 11 Vex . 56 For the equivalent circuit shown in Fig. 3-27(c), I2 = −

RTh =

Vex . Iex

In terms of our solution, Iex = −I2 . Hence, RTh = −

Vex 56 = .  I2 11

“book” — 2015/5/4 — 7:08 — page 149 — #35

´ 3-6 THEVENIN AND NORTON EQUIVALENT CIRCUITS

149

Table 3-1: Properties of Th´evenin/Norton analysis techniques. To Determine

Method

Can Circuit Contain Dependent Sources?

Relationship

υTh υTh

Open-circuit υ Short-circuit i (if RTh is known)

Yes Yes

υTh = υoc υTh = RTh isc

RTh RTh RTh

Open/short Equivalent R External source

Yes No Yes

RTh = υoc / isc RTh = Req RTh = υex / iex

iN = υTh /RTh ; RN = RTh

and

'

Thevenin and Norton Equivalency RTh '

Thevenin equivalent circuit

υTh

RN = RTh .

a

+ _

(3.37b)

Table 3-1 provides a summary of the various methods available for finding the elements of the Th´evenin and Norton equivalent circuits.

b Concept Question 3-12: Why is the Th´evenin-equivalent

circuit method such a powerful tool when analyzing a complex circuit, such as that of a cell phone? (See )

Norton equivalent circuit iN = υTh /RTh RN = RTh

iN

a

3-6.8 Analyzing Cascaded Systems

b

Let us go back to the simple household circuit of Fig. 3-19(b) and redraw it in Fig. 3-29(a) as a series combination of blocks: the voltage source consisting of Vs and associated resistance Rs , the fan, the lamp, and the DVM. We intend to use the circuit to demonstrate how the Th´evenin technique is used in practice to analyze much more complicated circuits. Our goal is to determine the voltage measured by the DVM.

RN

Figure 3-28: Equivalence between Th´evenin and Norton equivalent circuits, consistent with the source transformation method of Section 2-3.4.

3-6.7

Blocks 1 and 2

Norton’s Theorem

A corollary of Th´evenin’s theorem, Norton’s theorem states that a linear circuit can be represented at its output terminals by an equivalent circuit composed of a parallel combination of a current source iN and a resistor RN . Application of source transformation (Section 2-3.4) on the Th´evenin equivalent circuit shown in Fig. 3-28 leads to the straightforward conclusion that iN and RN of the Norton equivalent circuit are given by υTh iN = RTh

(3.37a)

We start with the combination of the first two blocks, namely the source and the fan, after disconnecting everything to the right of terminals (c, d) from the circuit. The Th´evenin voltage between terminals (c, d) in Fig. 3-29(b) is labeled Vcd and is given by Vcd =

Vs Rfan 110 × 100 = = 100 V. Rs + Rfan 10 + 100

The Th´evenin resistance of the circuit at terminals (c, d) in Fig. 3-29(b) is the parallel combination of Rs and Rfan : Rcd = Rs  Rfan = 10  100 = 9.09 .

“book” — 2015/5/4 — 7:08 — page 150 — #36

150

CHAPTER 3 ANALYSIS TECHNIQUES

a Vs

+ _

Rs = 10 Ω 110 V

c Rfan

b

Source block

e

Fan block

+

Rlamp 100 Ω

100 Ω d

f

Lamp block

_

DVM block

(a) Total circuit

a Vs

+ _

Rs = 10 Ω 110 V

Rfan b

Source block

Rcd = 9.09 Ω c

c

+ _

100 Ω Vcd

Fan block

Vcd

+ _

100 V

d

d

(b) Thévenin equivalent of source/fan combination

9.09 Ω Vcd

+ _

100 V

Re f = 8.33 Ω e

e

c Rlamp 100 Ω

d

+ _

Ve f

Ve f

+ _

91.67 V f

f Thévenin equivalent of source/fan combination

Thévenin equivalent of source/fan/lamp combination

(c) Thévenin equivalent of first three blocks

RTh = 8.33 Ω e + VTh = 91.67 V _

+ _

VDVM

DVM 2 MΩ

f (d) Final equivalent circuit

Figure 3-29: Repeated application of Th´evenin-equivalent circuit technique.

Blocks 1, 2, and 3

Blocks 1–4

Next, we repeat the Th´evenin technique at terminals (e, f ) by combining the lamp with the two earlier blocks. The Th´evenin voltage at terminals (e, f ) in Fig. 3-29(c) is labeled Vef and is given by 100 × 100 Vef = = 91.67 V, 9.09 + 100 and the Th´evenin resistance, Ref , is

In part (d) of Fig. 3-29, we show the Th´evenin equivalent of all blocks to the left of the DVM connected to the DVM at terminals (e, f ). Voltage division leads to

Ref = 100  9.09 = 8.33 .

This is the same answer we would have obtained had we analyzed the entire circuit at once using KCL/KVL. For

VDVM =

91.67 × 2 × 106 ≈ 91.67 V. 8.33 + 2 × 106

“book” — 2015/5/4 — 7:08 — page 151 — #37

3-7

COMPARISON OF ANALYSIS METHODS

151

this simple circuit, the multiple application of the Th´evenin equivalent technique is obviously unwarranted, but when dealing with complex circuits comprising multiple subsections, the Th´evenin technique is not only desirable, but also the only practical way to analyze and design circuits. Concept Question 3-13: Section 3-6 offers three

different approaches for finding RTh. Which ones apply to circuits containing dependent sources? (See )

Exercise 3-13: Find the Norton equivalent at terminals (a, b) of the circuit in Fig. E3.13.

3I a I 2A



10 Ω

b

Figure E3.13 Answer:

3-8

Maximum Power Transfer

Suppose an active linear circuit is connected to a passive linear circuit, as depicted by Fig. 3-30(a). An active circuit contains at least one independent source, whereas a passive circuit may contain dependent sources, but no independent sources. For convenience, we shall refer to them as the source and load circuits, respectively. For certain applications, it is desirable to maximize the magnitude of the current iL that flows from the source circuit to the load circuit, while other applications may call for maximizing the voltage υL at the input to the load circuit, or maximizing the power pL that gets transferred from the source to the load. Given a specified source circuit, how, then, does one approach the design of the load circuit so as to achieve these different goals? The solution to the problem posed by our question is facilitated by the equivalence offered by Th´evenin’s theorem. We demonstrated in the preceding section that any active, linear circuit always can be represented by an equivalent circuit composed of a Th´evenin voltage υTh connected in series with a Th´evenin resistance RTh . In the case of the passive load circuit, its equivalent circuit consists of only a Th´evenin resistance. To avoid confusion between the two circuits, we denote υTh and RTh of the source circuit as υs and Rs , and we denote RTh of the load circuit as RL , as shown in Fig. 3-30(b). The current iL

a 0.5 A

a



(See

+ υ −L

Active circuit

b

iL Passive circuit

b

)

Source circuit

(a) Source and load circuits

3-7 Comparison of Analysis Methods In this and the preceding chapter, we presented several different methods for analyzing electric circuits. Which method is best? Which one is the easiest to implement and why? The answer depends on the circuit configuration and the intended application. Table 3-2 provides a summary of the key attributes of the three circuit-analysis laws (Ohm’s law, KCL, and KVL) and the analysis methods covered thus far. If the circuit contains no dependent sources and the goal is to determine the currents and voltages in the circuit, the two by-inspection methods provide a straightforward solution approach. When dependent sources are present, the node voltage and mesh current methods are always applicable. For cascaded circuits, the Th´evenin (and Norton) equivalent-circuit technique is invariably the preferred choice.

Load circuit

Rs υs

+ _

a

+ υL



iL RL

b (b) Replacing source and load circuits with their Thévenin equivalents Figure 3-30: To analyze the transfer of voltage, current, and power from the source circuit to the load circuit, we first replace them with their Th´evenin equivalents.

“book” — 2015/5/4 — 7:08 — page 152 — #38

152

CHAPTER 3 ANALYSIS TECHNIQUES

Table 3-2: Summary of circuit analysis methods. Method

Common Use

Ohm’s law

Relates V , I , R. Used with all other methods to convert V ⇔ I .

R, G in series and 

Combine to simplify circuits. R in series adds, and is most often used. G in  adds, so may be used when much of the circuit is in parallel.

Y- or -T

Convert resistive networks that are not in series or in  into forms that can often be combined in series or in . Also simplifies analysis of bridge circuits.

Voltage/current dividers

Common circuit configurations used for many applications, as well as handy analysis tools. Dividers can also be used as combiners when used “backwards.”

Kirchhoff’s laws (KVL/KCL)

Solve for branch currents. Often used to derive other methods.

Node-voltage method

Solves for node voltages. Probably the most commonly used method because (1) node voltages are easy to measure, and (2) there are usually fewer nodes than branches and therefore fewer unknowns (smaller matrix) than KVL/KCL.

Mesh-current method

Solves for mesh currents. Fewer unknowns than KVL/KCL, approximately the same number of unknowns as node voltage method. Less commonly used, because mesh currents seem less intuitive, but useful when combining additional blocks in cascade.

Node-voltage by-inspection method

Quick, simplified way of analyzing circuits. Very commonly used for quick analysis in practice. Limited to circuits containing only independent current sources.

Mesh-current by-inspection method

Quick, simplified way of analyzing circuits. Very commonly used for quick analysis in practice. Limited to circuits containing only independent voltage sources.

Superposition

Simplifies circuits with multiple sources. Commonly used for both calculation and measurement.

Source transformation

Simplifies circuits with multiple sources. Commonly used for both calculation/design and measurement/test applications.

Th´evenin and Norton equivalents

Very often used to simplify circuits in both calculation and measurement applications. Also used to analyze cascaded systems. Th´evenin is the more commonly used form, but Norton is often handy for analyzing parallel circuits. Source transformation allows easy conversion between Th´evenin and Norton.

Input/output resistance (Rin /Rout )

Commonly used to evaluate when cascaded circuits can be analyzed individually or when full circuit analysis or a buffer is needed.

and associated voltage υL are given by Ohm’s law as iL =

υs , Rs + R L

and by voltage division: υL =

υ s RL . Rs + R L

“book” — 2015/5/4 — 7:08 — page 153 — #39

3-8

MAXIMUM POWER TRANSFER

153

 This equality is referred to as matching the source to the load. 

pL Maximum power when RL = Rs

pmax

The proof of Eq. (3.42) is given in Example 3-14. Use of RL = Rs in Eq. (3.41) leads to

RL

0 Rs

pmax =

υs2 RL υs2 = , (RL + RL )2 4RL

(3.43)

Figure 3-31: Variation of power pL dissipated in the load RL , as a function of RL . which represents 50 percent of the total power generated by the equivalent input source υs . The other 50 percent is dissipated in Rs . If the source-circuit parameters υs and Rs are fixed and the intent is to transfer maximum current to the load circuit, then RL should be zero (short circuit). For a real circuit with a functional purpose, the circuit will need to receive some energy in order to function. Hence, RL cannot be exactly zero, but it can be made to be very small in comparison with Rs . Thus, to maximize current transfer, the load circuit should be designed such that RL Rs

(maximum current transfer).

(3.39)

Based on Eq. (3.38), the opposite is true for maximum voltage transfer, namely RL Rs

(maximum voltage transfer).

(3.40)

The situation for power transfer calls for maximizing the product of iL and υL , υs2 RL pL = iL υL = . (Rs + RL )2

(maximum power transfer).

Prove that pL , as given by Eq. (3.41), is at a maximum when RL = Rs . Solution: To find the value of RL at which the expression for pL is at a maximum, we differentiate the expression with respect to RL and then set the result equal to zero. That is, dpL d = dRL dRL = υs2



υs2 RL (Rs + RL )2



1 2RL = 0. − (Rs + RL )2 (Rs + RL )3

A few simple steps of algebra lead to RL = Rs .

(3.41)

The expression given by Eq. (3.41) is a nonlinear function of RL . The power pL goes to zero as RL approaches either end of its range (0 and ∞), as illustrated by the plot in Fig. 3-31, and it is at a maximum when RL = Rs

Example 3-14: Maximum Power Transfer

(3.42)

Concept Question 3-14: Under what conditions is the power transferred from a power source to a load resistor a maximum? When is the voltage a maximum? When is the current a maximum? (See )

Concept Question 3-15: Of the power generated by an

input circuit, what is the maximum fraction that can be transferred to an external load? (See )

“book” — 2015/5/4 — 7:08 — page 154 — #40

154

TECHNOLOGY BRIEF 8: DIGITAL AND ANALOG

Technology Brief 8 Digital and Analog Most of electrical engineering depends on the manipulation of voltages and currents. The real world interfaces with our circuits through sensors (such as the resistive sensors in Technology Brief 4), and we interface back to the real world through user interfaces (such as turning on an LED in Technology Brief 5). In between these transducers are circuits! In the physical world, most signals of interest are analog signals; that is, they vary continuously with time and can take on any value between their possible minimum and maximum values. When electrical sensors transduce these signals into changes in voltage or current, the electrical signals produced are thus also analog. Analog electrical signals can be transduced from sound (using a microphone), mechanical vibration (using a piezoelectric vibration sensor), light or images (using sensor arrays in a camera), temperature (using a thermistor), and many other sources. All of the circuits we have examined so far are analog circuits. The voltages (and currents) present in these circuits can take on any value between a maximum and a minimum (typically set by the power source). By contrast, a digital signal can only assume a few discrete values. Most digital systems are binary, which is to say they can only assume two such values, usually called “0” and “1” (alternatively, “on” and “off”). The exact voltages which represent the two logic states depend on the type of digital logic used; for example, many modern digital processors represent “0” with 0 V and “1” with 1.2 V. Because any single digital line can only assume two values, many such lines can be used to represent a range of numbers. Consider, for example, Table TT8-1: three digital lines (or bits) are used to encode 8 different numbers within a given range. In the same way that

base-10 numbers can encode 10N different values with N discrete numbers in the range 0–9 (e.g., two base 10 numbers can encode 0–99), 2N different values can be encoded by N binary bits. Eight such bits make up a byte (e.g., the value 01101111 is a byte). Two bytes (16 bits) are a word. Standard encoding schemes exist for representing commonly used data. For example, letters, carriage returns, and other typographics can be represented using the 7-bit American Standard Code for Information Interchange (ASCII, pronounced “ask-ee”). Table TT8-2 gives these codes for capital letters. Many such standards exist (ranging from the data encoding format for, say, Blu-ray data to data transmission across ATM networks). When representing floating point numbers (such as −2.3), the computer must encode the sign (−1), the number and the exponent. The precision to which a number can be represented depends on how many bits are used. Four words (32 bits) are considered single precision, and 64 bits are double precision. The first bit is the sign (1 = negative), and the next 8 bits are the exponent. The remaining 23 bits (single precision) or 55 bits (double precision) are used to represent the number. This means that the floating point representation of the number has a certain predictable round-off error, and when the computer adds, subtracts, multiplies, etc., this error is also present in the calculations. Usually it is too small to be noticed, but in some cases (2 − 1.9999 . . . = 0) it can cause unexpected problems in computer programs. We commonly convert back and forth between analog and digital voltages. Almost all analog signals are converted to digital signals for storage (e.g., images), wireless transmission (your voice in a cell phone call), and performing mathematical functions (in your calculator). This is done with an analog-to-digital converter (ADC). Sometimes the digital signal must be converted back to

Table TT8-1: Three-bit counting scheme. Bits 000 001 010 011 100 101 110 111

Number = = = = = = = =

0 1 2 3 4 5 6 7

“book” — 2015/5/4 — 7:08 — page 155 — #41

TECHNOLOGY BRIEF 8: DIGITAL AND ANALOG

155

Table TT8-2: ASCII characters for capital letters. 0100 0001 = A 0100 0010 = B 0100 0011 = C 0100 0100 = D 0100 0101 = E 0100 0110 = F 0100 0111 = G 0100 1000 = H 0100 1001 = I 0100 1010 = J 0100 1011 = K 0100 1100 = L 0100 1101 = M

analog (so your friend can hear your voice on his cell phone). This is done with a digital-to-analog converter (DAC). The analog voltage in Fig. TF8-1 can be converted to digital using an ADC to sample it, find the closest step that matches the signal, and convert the value of that step to a digital value. The number of steps (controlled by the number of bits in the ADC) controls the precision of the ADC. Figure TF8-1 shows a very coarse 3-bit ADC that can represent 8 levels. The difference between the actual analog signal and the level that can be represented with the ADC is called the quantization error.

Analog

0100 1110 = N 0101 1111 = O 0101 0000 = P 0101 0001 = Q 0101 0010 = R 0101 0011 = S 0101 0100 = T 0101 0101 = U 0101 0110 = V 0101 0111 = W 0101 1000 = X 0101 1001 = Y 0101 1010 = Z

One of the strengths of digital representations of data is that manipulations of this data (mathematical operations, storage, etc.) can be carried out efficiently with switching networks.These are circuits of components wherein each component can only produce one of two voltage values. Transistors, in particular MOSFETS (see Chapter 4), are particularly well-suited to act as switches in these circuits; modern integrated circuits contain on the order of a billion MOSFETS arranged into circuits to manipulate digital data. Importantly, most modern integrated circuits contain both analog and digital circuits and are known as mixedsignal circuits (see Section 13-9). Using built-in ADC and

Digital

10 V = 111 8.0 V = 110 6.8 V = 101 5.7 V = 100 4.3 V = 011 2.8 V = 010 1.4 V = 001 0 V = 000

Time

Figure TF8-1: Three-bit digital representation of a continuous signal.

“book” — 2015/5/4 — 7:08 — page 156 — #42

156

TECHNOLOGY BRIEF 8: DIGITAL AND ANALOG

A AND B

A OR B

Input Output A B A and B 0 0 0 0 1 0 1 0 0 1 1 1

Input Output A B A or B 0 0 0 0 1 1 1 0 1 1 1 1

NOT

Input Output A Not A 0 1 1 0

A

Input Output A B A xor B 0 0 0 0 1 1 1 0 1 1 1 0

A XOR B

Figure TF8-2: Logic gates.

DAC circuits, data is moved from the analog to the digital domain within a single chip. But sometimes we use only a few gates for simple control operations or prototyping. Each gate takes two digital signals (which can be either a 0 or 1) as input, and outputs a different digital signal (based on these inputs). Figure TF8-2 shows a few of these common logic gates. An AND gate outputs a 1 if both input A AND input B are 1. An OR gate outputs a 1 if either input A OR input B are 1. A NOT gate outputs a 1 if input A is NOT a 1; i.e., it inverts the input. An exclusive OR gate, called an XOR gate, outputs a 1 if one and only one of input A OR input B is 1. One way to prototype with logic gates is to use a chip that plugs into your protoboard (see Appendix F). Figure TF8-3 shows an example of a quad AND package. Each pin on the chip is numbered 1–14 and plugs into a separate node (row) on the protoboard. Logic gates are active devices, which means they require an external power supply, so Vcc is plugged into pin 14, and GND into pin 7. Interfacing from the real world to a computer most often involves an analog sensor (such as a thermistor for measuring temperature), a level-shifter (amplifier or deamplifier or comparator that converts the analog output

14 Vcc

13

12

11

10

9

8

GND 1

2

3

4

5

6

7

7408 Quad 2 Input AND Figure TF8-3: Quad AND package.

voltage to digital levels), and then a logic circuit to act upon the output (turn a switch to a heater on or off, for instance). When interfacing back to the real world, the digital signal may be applied in digital form, or may need to be converted back to an analog signal (to drive speakers for voice and music, or precision control of an engine air intake, for example).

“book” — 2015/5/4 — 7:08 — page 157 — #43

3-8

MAXIMUM POWER TRANSFER

157

Exercise 3-14: The bridge circuit of Fig. E3.14 is connected to a load RL between terminals (a, b). Choose RL such that maximum power is delivered to RL . If R = 3 , how much power is delivered to RL ?

24 V

R

+ _

5Ω 16 V



+ _

4Ω a

2R

RL

b



a RL b



(a) Original circuit 2R

V1

R 5Ω

Figure E3.14 Answer: RL = 4R/3 = 4 , pmax = 4 W. (See

16 V )



+ _

Va

a

4Ω b + Voc _

I1

I2 4Ω

Example 3-15: Bridge Circuit

Vb



(b) Open-circuit voltage

In the bridge circuit shown in Fig. 3-32(a), choose RL so that the power delivered to it is a maximum. How much power will that be? Solution: After temporarily removing RL from the circuit, we proceed to find the Th´evenin equivalent of the circuit at terminals (a, b). Open-Circuit Voltage: In the circuit shown in Fig. 3-32(b), we designate the bottom node of the bridge as ground and the top node as voltage V1 . Application of KCL at node V1 gives



16 V

+ _

Voltage division gives

Isc

I3



4Ω b



(c) Short-circuit current RL

a

b

+_ 2.88 Ω

V1 = 6 V.

a

I1

V1 V1 V1 − 16 + + = 0, 5 2+4 2+4 which leads to

I2



2V

(d) Thévenin equivalent circuit Figure 3-32: Evolution of the circuit of Example 3-15.





4 V1 = 4 V, 2+4   2 Vb = V1 = 2 V. 2+4

Va =

Hence, VTh = Voc = Va − Vb = 4 − 2 = 2 V.

Short-Circuit Current: In the circuit configuration shown in Fig. 3-32(c), terminals (a, b) are connected by a short circuit. Application of the mesh-analysis by-inspection method (Section 3-4.2) leads to the matrix equation ⎡ ⎤⎡ ⎤ ⎡ ⎤ 16 11 −2 −4 I1 ⎣−2 6 0 ⎦ ⎣I2 ⎦ = ⎣ 0 ⎦ . 0 I3 −4 0 6

“book” — 2015/5/4 — 7:08 — page 158 — #44

158

CHAPTER 3 ANALYSIS TECHNIQUES

Matrix inversion by MATLAB or MathScript yields I1 =

96 A, 46

I2 =

32 A, 46

I3 =

pnp

64 A. 46

and RTh

32 64 32 − = = 0.7, 46 46 46

B

VCE

+ VBE IE

_

_

E

E

Configuration

Schematic symbol

(a) pnp transistor

npn

C

_ VBC

υ2 (2)2 = s = = 0.35 W. 4RL 4 × 2.88

n B

With the exception of the SPDT switch, all of the elements we have discussed thus far have been two-terminal devices, each characterized by a single i–υ relationship. These include resistors, voltage and current sources, as well as the pn-junction diode of Section 2-6.2. The potentiometer (Fig. 2-3(b)) may appear to be like a three-terminal device, but in reality it is no more than two resistors—each with its own pair of terminals. This section introduces a true three-terminal device, the bipolar junction transistor (BJT). The BJT is a three-layer semiconductor structure commonly made of silicon. Other compounds sometimes are used for special-purpose applications (such as for operation at microwave and optical frequencies), but for the present, we will limit our examination to silicon-based transistors and their uses in dc circuits. The three terminals of a BJT are called the emitter, collector, and base, and each is made of either a p-type (silicon with positive charge carriers) or n-type (silicon with negative charge carriers) semiconductor material. The emitter and collector are made of the same material—either p-type or n-type—and the base is made of the other material. Thus, the BJT can be constructed to have either a pnp configuration or an npn configuration, as shown in the diagrams of Fig. 3-33. The geometries and fabrication details of real transistors are

B

n

RL = RTh = 2.88 ,

3-9 Application Note: Bipolar Junction Transistor (BJT)

+

IC

+ IB

Conducting connector

The Th´evenin equivalent circuit is shown in Fig. 3-32(d). Power transfer to RL is a maximum when

pmax

VBC

C

p

Voc 2 = = 2.88 . = Isc 0.7

and according to Eq. (3.45),

_

Conducting connector

p

Hence, the short-circuit current is Isc = I3 − I2 =

C

C

+

IC

+ IB

p

B

VCE

+

n VBE IE

_

_ E

E

Configuration

Schematic symbol

(b) npn transistor Figure 3-33: Configurations and symbols for (a) pnp and (b) npn transistors.

far more elaborate than the simple diagrams suggest, but the basic idea that the BJT consists of three alternating layers of p- and n-type material is quite sufficient from the standpoint of its external electrical behavior. Figure 3-33 also shows schematic symbols used for the pnp and npn transistors. The center terminal is always the base. One of the three leads includes an arrow. The lead containing the arrow identifies the emitter terminal and whether the transistor is a pnp or npn. The arrow always points towards an n-type material, so in the pnp transistor, the arrow points towards the base, whereas in an npn transistor, the arrow points away from the base.

“book” — 2015/5/4 — 7:08 — page 159 — #45

3-9 APPLICATION NOTE: BIPOLAR JUNCTION TRANSISTOR (BJT)

IB

Example 3-16: BJT Amplifier Circuit

IC

B

C

Apply the equivalent-circuit model with VBE ≈ 0.7 V and β = 200 to determine IB , IC , and VCE in the circuit of Fig. 3-35(a). Assume that VBB = 2 V, VCC = 10 V, RB = 26 k, and RC = 200 .

βIB

VBE

159

IE

Solution: Upon replacing the npn transistor with its equivalent circuit, we end up with the circuit shown in Fig. 3-35(b). In the left-hand loop, KVL gives

E Figure 3-34: dc equivalent model for the npn transistor. The equivalent dc source VBE ≈ 0.7 V.

−VBB + RB IB + VBE = 0, which leads to IB =

The directions of the terminal currents shown in Fig. 3-33 are defined such that the base and collector currents IB and IC , respectively, flow into the transistor, and the emitter current IE flows out of it. KCL requires that

VBB − VBE 2 − 0.7 = = 5 × 10−5 A = 50 μA. RB 26 × 103

Given that β = 200, IC = βIB = 200 × 50 × 10−6 = 10 mA and

IE = IB + IC .

(3.44)

Under normal operating conditions, IE has the largest magnitude of the three currents, and IB is much smaller than either IC or IE . The transistor can operate under both dc and ac conditions, but we will limit our present discussion to dc circuits. For simplicity, we will consider only the npn common-emitter configuration. Accordingly, we can describe the operation of the npn transistor by the dc equivalent model shown in Fig. 3-34. The circuit contains a constant dc voltage source VBE and a dependent current-controlled current source that relates IC to IB by IC = βIB ,

VCE = VCC − IC RC = 10 − 10−2 × 200 = 8 V, which is a 4-times amplification of source VBB .

C RB

+

IC

RC

VBB

VCC

VCE

_

E (a) Transistor circuit

(3.45)

where β is a transistor parameter called the common-emitter current gain. This is a perfect example of how a nonlinear element can be modeled in terms of a linear circuit containing a dependent source. Under normal operation, VBE ≈ 0.7 V, and β may assume values in the range between 30 and 1000, depending on its specific design configuration. The VBE source in Fig. 3-34 models a built-in voltage drop that arises within the transistor at the interface of p-type and n-type regions; it is not a true independent source as it can never supply power. Transistors never supply power, they modify the flow of power through them in interesting and useful ways. To operate in its active mode, the transistor requires that certain dc voltages be applied at its base and collector terminals. We shall refer to these voltages as VBB and VCC , respectively.

B IB

RB

B IB

C

+

IC VBB

VBE

βIB

RC

VCC E (b) Equivalent circuit Figure 3-35: Circuit for Example 3-16.

VCE

_

“book” — 2015/5/4 — 7:08 — page 160 — #46

160

CHAPTER 3 ANALYSIS TECHNIQUES

Example 3-17: Digital-Inverter Circuit

+

RB = 20 kΩ Digital logic deals with two states, “0” and “1” (or equivalently “low” and “high”). A digital-inverter circuit provides one of the logic operations performed by a computer processor, namely to invert the state of an input bit from low to high or from high to low. Demonstrate that the transistor circuit shown in Fig. 3-36 functions as a digital inverter by plotting its output voltage Vout versus the input voltage Vin . A bit is assumed to be in state 0 (low) if its voltage is between 0 and 0.5 V and in state 1 (high) if its voltage is greater than 4 V. Assume that the equivalent model given by Fig. 3-34 is applicable (with β = 20) with the following qualifications: neither IB nor Vout can have negative values, so if the analysis using the equivalentcircuit model generates a negative value for either one of them, it should be replaced with zero. Solution: The equivalent circuit shown in Fig. 3-36(b) provides the following expressions: Vin − 0.7 , 20k IC = βIB = 200IB , IB =

(3.46)

RC = 1 kΩ

+ Vin

VCC = 5 V

_

(a)

_

Inverter circuit RB = 20 kΩ IB

IC

+

+ 1 kΩ

Vin

Vout

200IB

0.7 V

5V

_ (b)

_

Equivalent circuit

Vout (V)

(3.47)

State I II

5

and

4 Vout = VCC − IC RC .

Vout

Input Output Low High High Low

3

(3.48)

2 Combining the three equations leads to

1 Vout

βRC = VCC − (Vin −0.7) = 12−10Vin RB

(V). (3.49)

Since Vout is linearly related to Vin , the plot would be a straight line, as shown in Fig. 3-36(c), but we also have to incorporate the provisions that IB cannot be negative (which occurs when Vin < 0.7 V), and Vout cannot be negative (which occurs when Vin = 1.2 V). The resultant transfer function clearly satisfies the digital inverter requirements: Input: Low

Output: High

If Vin < 0.5 V

Vout = 5 V,

Input: High

Output: Low

If Vin > 1.2 V

Vout = 0.

II Vin (V)

0 0 0.7 1.2 2 (c)

I

3

4

5

Vout versus Vin

Figure 3-36: Circuit for Example 3-17.

Concept Question 3-16: How is the collector current

related to the base current in a BJT? (See

)

Concept Question 3-17: What is a digital inverter? How

are its input and output voltages related to one another? (See )

“book” — 2015/5/4 — 7:08 — page 161 — #47

NODAL ANALYSIS WITH MULTISIM

V(1) + _

161

100 Ω

2V

V(3)

50 Ω

50 Ω

+

V1 = 1 V V(4) _ +

3-10

Vx _

75 Ω

V(5)

0.1Vx

V(2) (a) Six-node circuit

(b) Multisim circuit and solution Figure 3-37: Circuit analysis with Multisim.

Exercise 3-15: Determine IB , Vout1 , and Vout2 in the

transistor circuit of Fig. E3.15, given that VBE = 0.7 V and β = 200.

5 kΩ

+

IB 200 Ω 8V

2V 100 Ω

+ _

Vout1

Figure E3.15 Answer: IB = 51.79 μA, Vout1 = 1.04 V,

Vout2 = 5.93 V. (See

)

Vout2

_

3-10

Nodal Analysis with Multisim

Multisim is a particularly useful tool for analyzing circuits with many nodes. Consider the six-node circuit shown in Fig. 3-37(a), in which the voltages and currents are designated in accordance with the Multisim notation system. In Multisim, V1 refers to the voltage of source 1 and V(1) refers to the voltage at node 1. Application of nodal analysis would generate five equations with five unknowns, V(1) to V(5), whose solution would require the use of matrix algebra or several steps of elimination of variables. [For this simple two-loop circuit, mesh analysis is much easier to apply, as it involves only two mesh equations and one auxiliary equation for the dependent current source, but the objective of the present section is to illustrate how Multisim can be used for circuits involving a large number of nodes.] When drawn in Multisim, the circuit appears in the form shown in Fig. 3-37(b). Application of either Measurement Probes or DC Operating Point Analysis generates the values of V(1) to V(5) listed in the inset of Fig. 3-37(b).

“book” — 2015/5/4 — 7:08 — page 162 — #48

162

CHAPTER 3 ANALYSIS TECHNIQUES

1V _

100 Ω

+

+

+ 2V _

50 Ω Vx _ 50 Ω

75 Ω

2 SPDT 1

3 kΩ

1A

I = 0.1Vx

(a) Circuit with SPDT switch

(b) Multisim configuration

Figure 3-38: (a) Circuit with a switch, and (b) its Multisim representation. For circuits containing more than four or five nodes, analyzing the circuit by hand becomes unwieldy. Moreover, some circuits may contain time-varying sources or elements. Consider, for example, the circuit in Fig. 3-38(a), which is a replica of the circuit in Fig. 3-37 except for the addition of an SPDT switch. [In Multisim, the switch can be toggled between positions 1 and 2 using the space bar on your computer.] When connected to position 1, the state of the circuit is identical to that in Fig. 3-37, but when the SPDT switch is moved to position 2, the new circuit configuration includes two additional elements and one extra node. The circuit drawn in Multisim is shown in Fig. 3-38(b). The SPDT is available in the Select a Component window under the Basic group in the SWITCH family. Measurement

Probes were added to nodes 4, 5, and 6. Using the Interactive Simulation feature of Multisim, the circuit can be analyzed in each of its two states by pressing F5 (or the

button or

switch) to start the simulation, and then toggling toggle the switch by pressing the space bar. This live-action switching capability is why this particular tool is known as Interactive Simulation. In the Multisim section of Chapter 2, we examined how the DC Operating Point Analysis tool can be used to determine differences between node voltages. In addition to basic subtraction, there are many operators that you can apply to variables (or combinations of variables) to obtain the desired

“book” — 2015/5/4 — 7:08 — page 163 — #49

3-10

NODAL ANALYSIS WITH MULTISIM

quantities. [See the Multisim Tutorial on the book website http://c3.eecs.umich.edu for a list of the basic operators]. We will now use variable manipulation in the DC Operating Point Analysis to calculate the power dissipated or supplied in each component in the circuit in Fig. 3-37(a). To calculate the power for each component, we need to know both the current through and voltage across each component. However, for many devices, Multisim can calculate the power automatically. Open up the DC Operating Point Analysis window. Notice that for the voltage sources and resistors, Multisim allows you to select to solve for the power, using the P() notation. You can also ask Multisim to solve for expressions which use the available variables. In the output tab enter equations via the Add Expression. . . button. We’ll enter an expression for the power across the controlled source this way using the expression V(5)*I(BI2). Click OK after entering any expressions. [Remember proper sign notation and current direction.] The equations for power should be Source V1: Source V2: Source I1: Resistor R1: Resistor R2: Resistor R3: Resistor R4:

(V(4)-V(3))*I(v1) (V(1)-V(2))*I(v2) -V(5)*I(v1) (V(3)-V(1))*I(v2) V(3)*I(v3) (V(5)-V(4))*I(v1) V(2)*I(v2)

163

(a) Multisim circuit of Fig. 3-32(a) ready for power calculations

(b) Selected variables for analysis visible in DC Operating Point Analysis window

Note: Remember that these variable names apply to the circuit shown in Fig. 3-39(a). If your circuit has a different numbering for nodes or voltage sources, your equations will differ in number accordingly. Once these equations are entered, the Selected Variables for Analysis field should resemble that in Fig. 3-39(b). To obtain the values, press the Simulate button. The results should agree with those shown in Fig. 3-39(c). Knowing how to write equations such as these in Multisim is very important, because many other Analyses which you will encounter later in the book utilize identical syntax to that used for the DC Operating Point Analysis. Concept Question 3-18: What is the difference between

the Measurement Probe tool and the DC Operating Point Analysis? (See )

(c) Output of simulations (remember that all values are in watts) Figure 3-39: Multisim procedure for calculating power

Exercise 3-16: Use Multisim to calculate the voltage at

node 3 in Fig. 3-38(b) when the SPDT switch is connected to position 2. Answer: (See

)

consumed (or generated) by the seven elements in the circuit of Fig. 3-37(a).

“book” — 2015/5/4 — 7:08 — page 164 — #50

164

CHAPTER 3 ANALYSIS TECHNIQUES

Summary Concepts • After designating one of the extraordinary nodes in a circuit as reference (ground), KCL at the remaining extraordinary nodes provides the requisite number of equations for determining the voltages at those nodes. • Two extraordinary nodes connected by a solitary voltage source constitute a supernode. The two nodes can be treated as a single node, augmented by an auxiliary relation specifying the voltage difference between them. • By assigning a mesh current to each independent loop, application of KVL leads to the requisite number of equations for determining the unknown mesh currents. • Two adjoining loops sharing a branch containing a solitary current source constitute a supermesh. The two loops can be treated as a single loop, augmented by an auxiliary relation specifying the relationship between their mesh currents.. • A circuit containing no dependent sources and only current sources can be analyzed by the node-voltage by-inspection method.

• Similarly, a circuit containing no dependent sources and only voltage sources can be analyzed the mesh-current by-inspection method. • Th´evenin’s (Norton’s) theorem states that a linear circuit can be represented by an equivalent circuit composed of a voltage source (current source) in series (in parallel) with a resistor. • Th´evenin and Norton equivalent circuits are powerful tools for analyzing and designing complex, cascaded circuits. • The power transferred by an input circuit to an external load is at a maximum when the load resistance is equal to the Th´evenin resistance of the input circuit. The fraction of the power thus transferred is 50 percent of the power supplied by the generator. • Multisim is a useful tool for simulating the behavior of a circuit and examining its sensitivity to specific variables of interest.

Mathematical and Physical Models

Mesh analysis by inspection

RI = Vt

Node-voltage method  of all current leaving a node = 0 [current entering a node is (−)]

Th´evenin equivalent circuit

υTh = υoc RTh = υoc / isc

Norton equivalent circuit

iN = isc RN = RTh

Maximum power transfer

RL = Rs

Mesh-current method  of all voltages around a loop = 0 [passive sign convention applied to mesh currents in clockwise direction] Nodal analysis by inspection

Important Terms active additivity property artificial sources base bipolar junction transistor (BJT) block

pmax =

GV = It

υs2 4RL

Provide definitions or explain the meaning of the following terms: bridge circuit by-inspection method cell-phone circuit collector common collector amplifier common-emitter amplifier

common-emitter current gain conductance matrix current mirror decoupled digital inverter emitter

“book” — 2015/5/4 — 7:08 — page 165 — #51

PROBLEMS

165

Important Terms (continued) extraordinary node homogeneity impedance independent linear circuit linear elements load circuit load impedance loading matching maximum power transfer mesh

mesh analysis by inspection mesh current nodal analysis by inspection node-voltage method Norton’s theorem npn configuration passive pnp configuration quasi-supernode resistance matrix scaling source circuit

*3.3 Use nodal analysis to determine the current Ix and amount of power supplied by the voltage source in the circuit of Fig. P3.3.

PROBLEMS Section 3-2: Node-Voltage Method *3.1 Apply nodal analysis to find the node voltage V in the circuit of Fig. P3.1. Use the information to determine the current I .

16 V

I



9A

V + _

source superposition source vector supermesh supernode superposition principle Th´evenin’s theorem Th´evenin’s voltage Th´evenin’s resistance uncoupled voltage vector



+ _ 40 V



Figure P3.3: Circuit for Problem 3.3.









+ _ 12 V

3.4 For the circuit in Fig. P3.4: (a) Apply nodal analysis to find node voltages V1 and V2 . (b) Determine the voltage VR and current I .

V1

1Ω Figure P3.1: Circuit for Problem 3.1.

16 V

+ _

3.2 Apply nodal analysis to determine Vx in the circuit of Fig. P3.2.

2Ω 1Ω

2Ω 3A

+ 4 Ω Vx _

Figure P3.2: Circuit for Problem 3.2.



Ix

Answer(s) available in Appendix G.



V2

1Ω + VR

_

I





Figure P3.4: Circuit for Problem 3.4.

*3.5 Apply nodal analysis to determine the voltage VR in the circuit of Fig. P3.5.

4Ω 12 V

+ _

+ VR

4Ω _



Figure P3.5: Circuit for Problem 3.5.

+ _

8V

“book” — 2015/5/4 — 7:08 — page 166 — #52

166

CHAPTER 3 ANALYSIS TECHNIQUES

3.6 Use the nodal-analysis method to find V1 and V2 in the circuit of Fig. P3.6, and then apply that to determine Ix .

V2

V1 4A



2A

V2



V1

Ix

12 Ω

4A Ix









+ _ 48 V

V3



3A Figure P3.9: Circuit for Problem 3.9.

Figure P3.6: Circuit for Problem 3.6.

3.10 The circuit in Fig. P3.10 contains a dependent current source. Determine the voltage Vx . *3.7

Find Ix in the circuit for Fig. P3.7.





10 Ω

+ 21 V _

Ix 10 Ω

2Ω + 6V _

5Ω _ + 10.5 V

Figure P3.7: Circuit for Problem 3.7.

3.8

I 8Ω



2Vx + _

Figure P3.11: Circuit for Problem 3.11.

3.12 The magnitude of the dependent current source in the circuit of Fig. P3.12 depends on the current Ix flowing through the 10  resistor. Determine Ix .

3A





+ 12 V _

(c) How much influence does the 4 A source have on the circuit to the left of the 3 A source?



+ Vx _

+ Vx _

(b) Determine the amount of power supplied by the voltage source.

+ _



*3.11 Determine the power supplied by the independent voltage source in the circuit of Fig. P3.11.

(a) Determine I .

6V

2Vx

Figure P3.10: Circuit for Problem 3.10.

For the circuit in Fig. P3.8:







4A

5Ω Ix

Figure P3.8: Circuit for Problem 3.8.

3.9 Apply nodal analysis to find node voltages V1 to V3 in the circuit of Fig. P3.9 and then determine Ix .

10 Ω

+ _ 12.3 V 4 Ω

20 Ω 2Ix



Figure P3.12: Circuit for Problems 3.12 and 3.13.

“book” — 2015/5/4 — 7:08 — page 167 — #53

PROBLEMS

167

*3.13 Repeat Problem 3.12 after replacing the 5  resistor in Fig. P3.12 with a short circuit.

4Ω 2I



I

1Ω + _ 8V

4V +_

0.2 Ω 0.5 Ω

_

+

3.14 Apply nodal analysis to find the current Ix in the circuit of Fig. P3.14.

6Ω 1Ω



+ Vx _

0.5 Ω

0.1 Ω + _ 2V

Ix

0.1 Ω

0.1 Ω

+ 3V _

Figure P3.17: Circuit for Problems 3.17 and 3.18.

3.18 Repeat Problem 3.17 after replacing the 2  resistor in Fig. P3.17 with a short circuit. 3.19

Figure P3.14: Circuit for Problem 3.14.

For the circuit shown in Fig. P3.19:

(a) Determine Req between terminals (a, b). *3.15 Use the supernode concept to find the current Ix in the circuit of Fig. P3.15.

(b) Determine the current I using the result of (a). (c) Apply nodal analysis to the original circuit to determine the node voltages and then use them to determine I . Compare the result with the answer of part (b).

2A 6V +

R

0.5 Ω

R

_

Ix

0.5 Ω

4A

0.5 Ω

R

I

3.16 Apply the supernode technique to determine Vx in the circuit of Fig. P3.16.

2 kΩ 6V +

+_

Req

b

R

R

V0

*3.20

For the circuit in Fig. P3.20, determine the current Ix .

6 kΩ 1Ω 0.1 Ω

5 kΩ + _ 10 V

4 kΩ

+ 4V _

0.2 Ω 1Ω

0.2 Ω Figure P3.16: Circuit for Problem 3.16.

*3.17

R

R

Figure P3.19: Circuit for Problem 3.19.

_

1 kΩ

+ Vx _

R

R

R a

Figure P3.15: Circuit for Problem 3.15.

R

Determine Vx in the circuit of Fig. P3.17.

Figure P3.20: Circuit for Problem 3.20.

Ix

0.1 Ω

“book” — 2015/5/4 — 7:08 — page 168 — #54

168

CHAPTER 3 ANALYSIS TECHNIQUES

3.21 Apply nodal analysis to determine Vx in the circuit of Fig. P3.21.



+ 2A



4Ω 2A





Vx

_



+ _

2V





+ 3Ω



1A



Vx

_

Figure P3.24: Circuit for Problem 3.24.

3.25 Apply nodal analysis to determine Va , Vb , and Vc in the circuit of Fig. P3.25.

Figure P3.21: Circuit for Problem 3.21.

3.22 Apply nodal analysis to determine VL in the circuit of Fig. P3.22.

2 kΩ

15 Ω 10 V Vb + _

Va

3 kΩ

2.5 Ω

1V

+

+ _

4 kΩ



1 kΩ + _

3A

_ +

25 V

+

V

_x

Vc 10 Ω

3Vx

2 kΩ

VL

_

+ _

50 V

7.5 Ω



Figure P3.22: Circuit for Problem 3.22.

*3.23 Apply nodal analysis to determine Vx in the circuit of Fig. P3.23.

Figure P3.25: Circuit for Problem 3.25.

Section 3-3: Mesh-Current Method

5V _

+ Vx _

*3.26 Apply mesh analysis to find the mesh currents in the circuit of Fig. P3.26. Use the information to determine the voltage V .

+

7 kΩ 2 kΩ

5 kΩ

2A

7V

+ _

3 kΩ 2Ω

Figure P3.23: Circuit for Problem 3.23.

3.24 Apply nodal analysis to determine Vx in the circuit of Fig. P3.24.

16 V

+ _

V

2Ω 4Ω

I1



I2

Figure P3.26: Circuit for Problem 3.26.

+ _ 12 V

“book” — 2015/5/4 — 7:08 — page 169 — #55

PROBLEMS

169

3.27 Use mesh analysis to determine the amount of power supplied by the voltage source in the circuit of Fig. P3.27.

4A 3Ω

8Ω 2Ω

9A

+ _ 40 V



6Ω 2Ω





+ _ 48 V

Figure P3.27: Circuit for Problem 3.27. Figure P3.31: Circuit for Problem 3.31.

*3.28 Determine V in the circuit of Fig. P3.28 using mesh analysis.

V

4Ω + 12 V _

*3.32 Use the supermesh concept to solve for Vx in the circuit of Fig. P3.32.

4Ω 2Ω

+ _ 8V





1Ω Figure P3.28: Circuit for Problem 3.28.

+ 16 V _

3.33 Use the supermesh concept to solve for Ix in the circuit of Fig. P3.33.

1Ω I 1Ω



*3.30 Apply mesh analysis to find Ix in the circuit of Fig. P3.30.

+ 21 V _

Ix 10 Ω

10 Ω



2A

5Ω 5Ω

3.31 Apply mesh analysis to determine the amount of power supplied by the voltage source in Fig. P3.31.

4A



3A

Figure P3.33: Circuit for Problem 3.33.

3.34 Apply mesh analysis to the circuit in Fig. P3.34 to determine Vx .

_ + 10.5 V

Figure P3.30: Circuit for Problem 3.30.

Ix

12 Ω



Figure P3.29: Circuit for Problem 3.29.



+ 4 Ω Vx _

Figure P3.32: Circuit for Problem 3.32.

3.29 Apply mesh analysis to find I in the circuit of Fig. P3.29.



3A

2Ω + 6V _



2Vx

Figure P3.34: Circuit for Problem 3.34.



+ Vx _

“book” — 2015/5/4 — 7:08 — page 170 — #56

170

CHAPTER 3 ANALYSIS TECHNIQUES

3.35 Determine the amount of power supplied by the independent voltage source in Fig. P3.35 by applying the meshanalysis method.

2A 6V + _

+ Vx _

Ix

0.5 Ω



+ 12 V _

0.5 Ω

2Vx + _



4A

0.5 Ω

Figure P3.39: Circuit for Problem 3.39. Figure P3.35: Circuit for Problem 3.35.

3.40 Use mesh analysis to find Ix in the circuit of Fig. P3.36.



4V +_

0.2 Ω 0.5 Ω 0.1 Ω + _ 2V

2I



0.5 Ω Ix

0.1 Ω

0.1 Ω

+ 3V _

I

1Ω + _ 8V

_



+

*3.36

Determine Vx in the circuit of Fig. P3.40.





+ Vx _

Figure P3.40: Circuit for Problems 3.40 and 3.42. Figure P3.36: Circuit for Problem 3.36.

3.37 The circuit in Fig. P3.37 includes a dependent current source. Apply mesh analysis to determine Ix .

3.41 Apply the supermesh technique to find Vx in the circuit of Fig. P3.41.

5Ω 10 Ω

+ _ 12.3 V 4 Ω

20 Ω

6V + 2Ix

_

Ix

2 kΩ

2Ω 1 kΩ

+ Vx _

6 kΩ 5 kΩ 4 kΩ 2 mA

Figure P3.37: Circuit for Problems 3.37 and 3.38. Figure P3.41: Circuit for Problem 3.41.

3.38 Repeat Problem 3.37 after replacing the 5  resistor in Fig. P3.37 with a short circuit. *3.39 Apply mesh analysis to the circuit of Fig. P3.39 to determine Ix .

*3.42 Repeat Problem 3.40 after replacing the 2  resistor in Fig. P3.40 with a short circuit.

“book” — 2015/5/4 — 7:08 — page 171 — #57

PROBLEMS

171

3.43 Apply mesh analysis to the circuit of Fig. P3.43 to find Ix .

3.46 Simplify the circuit in Fig. P3.46 as much as possible using source transformation and resistance combining, and then apply mesh analysis to determine Ix .

1Ω 0.1 Ω + 4V _

0.2 Ω Ix



0.2 Ω

6Ω 12 V



+ _





Ix



6Ω 1Ω 4Ω

0.1 Ω



Figure P3.46: Circuit for Problem 3.46. Figure P3.43: Circuit for Problem 3.43.

3.44

Determine I0 in Fig. P3.44 through mesh analysis.

3.47 Apply mesh analysis to determine I0 in the circuit in Fig. P3.47.

2Ω 4Ω



_ 10 V +



I0





2V _

4Ix





+

+ 10 V _

Ix

4Ω I0

4Ω 6Ω



Figure P3.44: Circuit for Problem 3.44. Figure P3.47: Circuit for Problem 3.47.

*3.45 Use an analysis method of your choice to determine I0 in the circuit of Fig. P3.45. *3.48 Apply mesh analysis to determine Ix in the circuit in Fig. P3.48.

+ 12 V _

5Ω 10 Ω

10 Ω

15 V +_





10 Ω

Ix

I0 10 Ω

10 Ω

2.5 A

5Ω Figure P3.45: Circuit for Problem 3.45.

Figure P3.48: Circuit for Problem 3.48.



“book” — 2015/5/4 — 7:08 — page 172 — #58

172

CHAPTER 3 ANALYSIS TECHNIQUES

3.49 Apply mesh analysis to determine Ix in the circuit in Fig. P3.49.

(c) The values of how many of those mesh currents can be determined immediately from the circuit? (d) Apply mesh analysis to find I  .

Ix

10 Ω



4A 5Ω

1A

+ _

10 Ω

5Ω _ +

50 V

+_

15 Ω

5V

85 V



20 Ω 7A

10 V 5Ω 5Ω _



10 V + _

I′

+

10 Ω

12 Ω 10 Ω

10 Ω 6A



3A

+ _

40 V

Figure P3.49: Circuit for Problem 3.49.

7.5 Ω 1A 3.50 Apply mesh analysis to determine Vx in the circuit in Fig. P3.50.

Figure P3.51: Circuit for Problem 3.51.

Sections 3-4 and 3-5: By-Inspection and Superposition Methods

1A





3Ω +_

4Ω 2Ω

*3.52 Apply the by-inspection method to develop a node-voltage matrix equation for the circuit in Fig. P3.52 and then use MATLAB or MathScript software to solve for V1 and V2 .

2V 12 Ω



+ Vx _ 2Ω

V1 6Ω

2A

V2 6Ω

4A



3A

Figure P3.50: Circuit for Problem 3.50. Figure P3.52: Circuit for Problem 3.52.

3.51

Consider the circuit shown in Fig. P3.51.

(a) How many extraordinary nodes does it have? (b) How many independent meshes does it have?

3.53 Use the by-inspection method to establish a node-voltage matrix equation for the circuit in Fig. P3.53. Solve the matrix equation by MATLAB or MathScript software to find V1 to V4 .

“book” — 2015/5/4 — 7:08 — page 173 — #59

PROBLEMS

173

2Ω 1Ω V2

V1 2A

6Ω 7Ω



V3 9Ω

2 kΩ 4Ω



4 kΩ

V4 3A



1 kΩ

Figure P3.53: Circuit for Problem 3.53.

+ 21 V _

4 kΩ 2 mA

3.57 Use the by-inspection method to establish the meshcurrent matrix equation for the circuit in Fig. P3.57 and then solve the equation to determine Vout .

16 Ω

5Ω I1

5 kΩ

+ Vx _

Figure P3.56: Circuit for Problem 3.56.

3.54 Develop a mesh-current matrix equation for the circuit in Fig. P3.54 by applying the by-inspection method. Solve for I1 to I3 .



6 kΩ

I2

I3

10 Ω

10 Ω

5Ω _ + 4.2 V

8Ω + _ 538 V

4Ω 8Ω 4Ω



Figure P3.54: Circuit for Problem 3.54.

+

2Ω 1Ω

3.55 Find I0 in the circuit of Fig. P3.55 by developing a meshcurrent matrix equation and then solving it using MATLAB or MathScript software.

Vout

_

Figure P3.57: Circuit for Problem 3.57.

*3.58 Develop a node-voltage matrix equation for the circuit in Fig. P3.58. Solve it to determine I .

+ 12 V _

10 Ω

20 Ω

20 Ω



10 Ω

20 Ω 20 Ω

V1



2A 5Ω

I0 10 Ω

I

20 Ω

V3

V2 3Ω



Figure P3.55: Circuit for Problem 3.55. Figure P3.58: Circuit for Problem 3.58.

*3.56 Apply the by-inspection method to derive a node-voltage matrix equation for the circuit in Fig. P3.56 and then solve it using MATLAB or MathScript software to find Vx .

3.59 Determine the amount of power supplied by the voltage source in Fig. P3.59 by establishing and then solving the meshcurrent matrix equation of the circuit.

“book” — 2015/5/4 — 7:08 — page 174 — #60

174

CHAPTER 3 ANALYSIS TECHNIQUES



*3.62 Perform necessary source transformations and then use the mesh analysis by-inspection method to determine Vx in the circuit of Fig. P3.62.

1Ω 3Ω 4Ω

+ _ 8V







5Ω 2A



3Ω 4Ω



+ Vx _

Figure P3.59: Circuit for Problem 3.59.

7Ω 3.60 Determine the current Ix in the circuit of Fig. P3.60 by applying the source-superposition method. Call Ix the component of Ix due to the voltage source alone, and Ix the component due to the current source alone. Show that Ix = Ix + Ix is the same as the answer to Problem 3.9.

4A 3Ω



Ix

2Ω 2Ω

+ _

4Ω 48 V

Figure P3.62: Circuit for Problem 3.62.

3.63 Apply the source-superposition method to the circuit in Fig. P3.63 to determine: (a) Vx , the component of Vx due to the 1 A current source alone. (b) Vx , the component of Vx due to the 10 V voltage source alone. (c) Vx , the component of Vx due to the 3 A current source alone. (d) The total voltage Vx = Vx + Vx + Vx .

+ Vx _

3.61 Apply the source-superposition method to the circuit in Fig. P3.61 to determine:

1A

12 Ω

(a) Ix , the component of Ix due to the voltage source alone (b)

1A



Figure P3.60: Circuit for Problem 3.60.

Ix ,



+ _

18 Ω

15 Ω 10 V

+ _

10 Ω



3A



the component of Ix due to the current source alone

(c) The total current Ix = Ix + Ix

Figure P3.63: Circuit for Problem 3.63.

(d) P  , the power dissipated in the 4  resistor due to Ix

(e) P  , the power dissipated in the 4  resistor due to Ix (f) P , the power dissipated in the 4  resistor due to the total current I . Is P = P  + P  ? If not, why not?

9A



Ix

8Ω 4Ω

Figure P3.61: Circuit for Problem 3.61.

+ _ 40 V

Section 3-6: Th´evenin and Norton Equivalents *3.64 Find the Th´evenin equivalent circuit at terminals (a, b) for the circuit in Fig. P3.64.

2Ω 1Ω

2Ω 3A

3Ω 4Ω

Figure P3.64: Circuit for Problem 3.64.

a b

“book” — 2015/5/4 — 7:08 — page 175 — #61

PROBLEMS

175

3.65 Find the Th´evenin equivalent circuit at terminals (a, b) for the circuit in Fig. P3.65.

2.5 Ω a 3Ω 4A





b + _ 2V

d

a

+ _

+ _ 19 V



2Ix

I0

b

b

8V +_

4Ω d

4Ω 2Ω e

+ 12 V _

Figure P3.68: Circuit for Problems 3.68 through 3.70.

a

b c

0.25 Ω

_ 0.2I + 0

b

Figure P3.73: Circuit for Problem 3.73.

*3.74 Find the Norton equivalent circuit at terminals (a, b) of the circuit in Fig. P3.74.



I0

a



0.2 Ω 0.2 Ω

0.1 Ω

*3.68 Find the Th´evenin equivalent circuit at terminals (a, b) for the circuit in Fig. P3.68.

+ _ 6V



a

Figure P3.71: Circuit for Problems 3.71 and 3.72.

3.67 For the circuit in Fig. P3.66, find the Th´evenin equivalent circuit as seen by the 6  resistor connected between terminals (c, d) as if the 6  resistor is a load resistor connected to (but external to) the circuit. Determine the current flowing through that resistor.





20 Ω

10 Ω

48 V

Figure P3.66: Circuit for Problems 3.66 and 3.67.



10 Ω

3.73 Find the Norton equivalent circuit at terminals (a, b) for the circuit in Fig. P3.73.





3.71 Find the Th´evenin equivalent circuit at terminals (a, b) of the circuit in Fig. P3.71.

*3.72 Find the Norton equivalent circuit of the circuit in Fig. P3.71 after increasing the magnitude of the voltage source to 38 V.

4A 6Ω

3.70 Repeat Problem 3.68 for terminals (d, e) as seen by the 2  resistor between them (as if it were a load resistor external to the circuit).

Ix

3.66 The circuit in Fig. P3.66 is to be connected to a load resistor RL between terminals (a, b). (a) Find the Th´evenin equivalent circuit at terminals (a, b). (b) Choose RL so that the current flowing through it is 0.5 A.

c

Repeat Problem 3.68 for terminals (a, c).



Figure P3.65: Circuit for Problem 3.65.



3.69

a

3Ω + _ 15 V



1.2I0

Figure P3.74: Circuit for Problems 3.74 and 3.75.

b

“book” — 2015/5/4 — 7:08 — page 176 — #62

176

CHAPTER 3 ANALYSIS TECHNIQUES

3.75 Repeat Problem 3.74 after replacing the 6  resistor with an open circuit.

a

3.76 Find the Norton equivalent circuit at terminals (a, b) of the circuit in Fig. P3.76.

I0

0.2 Ω



a

−+ 2I0

0.2 Ω

0.1 Ω



I0



+ _ 0.2I0

4Ω b

b Figure P3.79: Circuit for Problem 3.79.

Figure P3.76: Circuit for Problems 3.76.

*3.77 Obtain the Th´evenin equivalent circuit at terminals (a, b) in Fig. P3.77.



b 1V

+ _



6Ω + _



5V

*3.80 Obtain the Th´evenin equivalent of the circuit in Fig. P3.80 at terminals (a, b).

a



4Ω 3Ω

Figure P3.77: Circuit for Problem 3.77.



3Ω 1Ω

3.78 Obtain the Th´evenin equivalent of the circuit to the left of terminals (a, b) in Fig. P3.78. Use your result to compute the power dissipated in the 0.4  load resistor.

a

b 2A

1A

Figure P3.80: Circuit for Problem 3.80.

a

3Ω 2Ω



Section 3-8: Maximum Power Transfer



1A + _

0.4 Ω 2V

3.81 What value of the load resistor RL will extract the maximum amount of power from the circuit in Fig. P3.81, and how much power will that be?

b 4Ω

3Ω Figure P3.78: Circuit for Problem 3.78.

3.79 Obtain the Th´evenin equivalent of the circuit in Fig. P3.79 at terminals (a, b).



4Ω 3A

6Ω 8Ω

Figure P3.81: Circuit for Problem 3.81.

a b

RL

“book” — 2015/5/4 — 7:08 — page 177 — #63

PROBLEMS

177

3.82 For the circuit in Fig. P3.82, choose the value of RL so that the power dissipated in it is a maximum.

+ _ 15 V

a 2 kΩ

Rs

3 kΩ

IL RL

6 kΩ

4 kΩ 2 mA

Figure P3.85: Circuit for Problem 3.85.

RL 6 kΩ

8 kΩ

b

Figure P3.82: Circuit for Problem 3.82.

*3.83 Determine the maximum power that can be extracted by the load resistor from the circuit in Fig. P3.83.

3.86 In the circuit shown in Fig. P3.86, a potentiometer is connected across the load resistor RL . The total resistance of the potentiometer is R = R1 + R2 = 5 k. (a) Obtain an expression for the power PL dissipated in RL for any value of R1 . (b) Plot PL versus R1 over the full range made possible by the potentiometer’s wiper.

2 kΩ 4 kΩ

2000Ix +_

+ _ 12 V

Ix

3 kΩ + _ 15 V

RL 1 kΩ

R

RL

6 kΩ

R1 R2

Figure P3.86: Circuit for Problem 3.86.

Figure P3.83: Circuit for Problem 3.83.

3.87 Determine the maximum power extractable from the circuit in Fig. P3.87 by the load resistor RL .

3.84 Figure P3.84 depicts a 0-to-10 k potentiometer as a variable load resistor RL connected to a circuit of an unknown architecture. When the wiper position on the potentiometer was adjusted such that RL = 1.2 k, the current through it was measured to be 3 mA, and when the wiper was lowered so that RL = 2 k, the current decreased to 2.5 mA. Determine the value of RL that would extract maximum power from the circuit.

I0

2 kΩ 2 kΩ

1 kΩ

RL

+ _ 200I0

Figure P3.87: Circuit for Problem 3.87.

Circuit

IL

a

} RL

b

3.88 In the circuit Fig. P3.88, what value of Rs would result in maximum power transfer to the 10  load resistor?

Figure P3.84: Circuit for Problem 3.84.

2A *3.85 The circuit shown in Fig. P3.85 is connected to a variable load RL through a resistor Rs . Choose Rs so that IL never exceeds 4 mA, regardless of the value of RL . Given that choice, what is the maximum power that RL can extract from the circuit?

Rs

RL

Figure P3.88: Circuit for Problem 3.88.

10 Ω

“book” — 2015/5/4 — 7:08 — page 178 — #64

178

CHAPTER 3 ANALYSIS TECHNIQUES

Section 3-9: Bipolar Junction Transistor *3.89 The two-transistor circuit in Fig. P3.89 is known as a current mirror. It is useful because the current I0 controls the current IREF regardless of external connections to the circuit. In other words, this circuit behaves like a current-controlled current source. Assume both transistors are the same size such that IB1 = IB2 . Find the relationship between I0 and IREF . (Hint: You do not need to know what is connected above or below the transistors. Nodal analysis will suffice.)

I0

IREF C1

C Iin

Rin

+ V0 _

C2

Transistor 1

3.91 The circuit in Fig. P3.91 is identical to the circuit in Fig. P3.90, except that the voltage source Vin is more realistic in that it has an associated resistance Rin . Find both the voltage gain (AV = Vout /Vin ) and the current gain (AI = Iout /Iin ). Assume Vin VBE .

(Power supply)

B E

+ _ Vin

Iout RL

Transistor 2 B

E1

Figure P3.89: A simple current mirror (Problem 3.89).

(Power supply)

RL + V0 _ (Power supply)

C

V0

_

3.92 The circuit in Fig. P3.92 is a BJT common-emitter amplifier. Find Vout as a function of Vin .

3.90 The circuit in Fig. P3.90 is a BJT common collector amplifier. Obtain expressions for both the voltage gain (AV = Vout /Vin ) and the current gain (AI = Iout /Iin ). Assume Vin VBE .

+ _

Vout

Figure P3.91: Circuit for Problem 3.91.

E2

Iin

+

B

+ Vout

_

Rs + _ Vin

E + _ Vin

Iout RL

Figure P3.90: Circuit for Problem 3.90.

+

Figure P3.92: Circuit for Problem 3.92.

Vout

_ *3.93 Obtain an expression for Vout in terms of Vin for the common emitter-amplifier circuit in Fig. P3.93. Assume Vin VBE .

“book” — 2015/5/4 — 7:08 — page 179 — #65

PROBLEMS

179

RL

+ V0 _ (Power supply)

+ Vx _

+ Vout

_



+ 12 V _

Rs

2Vx + _



Figure P3.97: Circuit for Problem 3.97.

+ _ Vin

3.98 Use the DC Operating Point Analysis in Multisim to find the power dissipated or supplied by each component in the circuit in Fig. P3.98 and show that the sum of all powers is zero.

RE

R1

Figure P3.93: Circuit for Problem 3.93.

2.5I 3.94 Using Multisim, draw the circuit in Fig. P3.94 and solve for voltages V1 and V2 .

R4

+

_

Section 3-10: Multisim Analysis

25 Ω R2

R3







R5 10 Ω I

R6 10 Ω

+ 10 V _

12 Ω V1

V2 6Ω

3A

Figure P3.98: Circuit for Problem 3.98.



4A

3A

Figure P3.94: Circuit for Problem 3.94.

3.95 The circuit in Problem 3.55 was solved using MATLAB or MathScript software. It can be solved just as easily using Multisim. Using Multisim, draw the circuit in Fig. P3.55 and solve for all node voltages and the current I0 . 3.96 Using Multisim, draw the circuit in Fig. P3.96 and solve for Vx .

3.99 Simulate the circuit found in Fig. P3.99 with a 10  resistor placed across the terminals (a, b). Then either by hand or by using tools in Multisim (see Multsim Demo 3.3), find the Th´evenin and Norton equivalent circuits and simulate both of those circuits in Multisim with 10  resistors across their output terminals. Show that the voltage drop across and current through the 10  load resistor is the same in all three simulations.

R1 50 Ω + _ 12 V

a I R2

R3 10 Ω 25 Ω

+ _ 2I

2Ω + 6V _



2Vx



+ Vx _

b

Figure P3.99: Circuit for Problem 3.99.

Potpourri Questions Figure P3.96: Circuit for Problem 3.96.

3.100 Why is it of interest to measure the conductivity of sea ice?

3.97 Use Multisim to draw the circuit in Fig. P3.97 and solve for Vx .

3.101 In integrated circuit fabrication, what is a wafer? A die? A chip?

“book” — 2015/5/4 — 7:08 — page 180 — #66

180

CHAPTER 3 ANALYSIS TECHNIQUES

3.102 How is lithography related to feature size in IC fabrication? Why are ICs fabricated under super-clean conditions? 3.103 What is a bit in a digital signal? A byte? A word? What does the acronym ASCII stand for?

R4

+ Isrc

V1 I1

_

R1 I2

R2 I3

R3 I4

_ + Vsrc

Figure m3.2 Circuit for Problem m3.2.

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials .] and videos are available on m3.1 Node-Voltage Method: Apply the node-voltage method to determine node voltages V1 to V4 for the circuit of Fig. m3.1. From these results determine which resistor dissipates the most power and which resistor dissipates the least power, and report these two values of power. Use these component values: Isrc1 = 3.79 mA, Isrc2 = 1.84 mA, Vsrc = 4.00 V, R1 = 3.3 k, R2 = 2.2 k, R3 = 1.0 k, and R4 = 4.7 k.

m3.3

Superposition: In the circuit of Fig. m3.3:

(a) Solve for Ia and Vb using nodal analysis. (b) Solve for Ia and Vb using superposition. Hint: Solve for Ia and Vb with one source on at a time. (c) Determine Ia and Vb using any method. Use these component values: I1 = 1.84 mA, V2 = 3.0 V, R1 = 1.0 k, R2 = 2.2 k, and R3 = 4.7 k.

+ Vb _ R3

Ia R2 R1

V1

I1 R4

V2

R2

R1 + _

V4

V2

R3 Isrc1

V3

+ _

Isrc2 Vsrc

Figure m3.1 Circuit for Problem m3.1.

m3.2 Mesh-Current Method: Apply the mesh-current method to determine mesh currents I1 to I4 in the circuit of Fig. m3.2. From these results determine V1 , the voltage across the current source. Use these component values: Isrc1 = 12.5 mA, Vsrc = 15 V, R1 = 5.6 k, R2 = 2.2 k, R3 = 3.3 k, and R4 = 4.7 k.

Figure m3.3 Circuit for Problem m3.3.

m3.4 Th´evenin Equivalents and Maximum Power Transfer: In the circuit of Fig. m3.4, find the Th´evenin equivalent of the circuit at terminals (a, b) as would be seen by a load resistor RL . Specifically: (a) Determine the open-circuit voltage Voc that appears at terminals (a, b). (b) Determine the short-circuit current Isc that flows through a wire connecting terminals (a, b) together.

“book” — 2015/5/4 — 7:08 — page 181 — #67

PROBLEMS

181

(c) Determine the Th´evenin resistance. (d) Determine the maximum power Pmax that could be delivered by this circuit. Use these component values: Vsrc = 10 V, R1 = 680 , R2 = 3.3 k, R3 = 4.7 k, and R4 = 1.0 k.

R1 Vsrc

+ _

(b) Add a short circuit between nodes 1 and 2, and then find the short circuit current between them. Use this information to calculate the Th´evenin resistance. (c) Turn off the 4 V and 8 V sources. Verify the Th´evenin resistance from part (b) by measuring the equivalent resistance between terminals 1 and 2 (using Multisim and myDAQ).

R3 a

R2

b

1

_

+ 2

+ _

V1 2V

R1

R3

1

R4

15 kΩ

R2

47 kΩ

4.7 kΩ

+ V2 _ 4V

1 kΩ

Figure m3.4 Circuit for Problem m3.4. Figure m3.6 Circuit for Problem m3.6. m3.5

Power Dissipation: For the circuit shown in Fig. m3.5:

(a) Find the combined total power generated by the two current sources analytically and with Multisim. Do not build this circuit (there is no myDAQ portion for part (a)). (b) Use source transformations to reduce the current sources in Fig. m2.5 into a single voltage source. Now, build this circuit and measure the total power dissipated by all four resistors. Hint: To create the voltage source, use the myDAQ arbitrary waveform generator. (c) Is the power found in part (a) the same as in part (b)?

m3.7 Power Dissipation with Current Source: Creating an ideal current source with the myDAQ requires a current regulator. For the myDAQ portion of this problem use the LM371 and a 220  resistor to create the current source in Fig. m3.7. (a) Determine the power generated by the current source. For the myDAQ portion of this problem, be sure to measure the current through the LM371 regulator. (b) Determine the total power dissipated by all other circuit elements. Compare your answer to the result obtained in part (a).

R1 3.3 kΩ I1

0.4 mA

R4 22 kΩ R3

R2 1 kΩ

I2 0.8 mA

R4 R1

3.3 kΩ

1 kΩ

4.7 kΩ

3.3 kΩ Figure m3.5 Circuit for Problem m3.5.

I1

5.68 mA

R2

R3 1 kΩ

m3.6 Th´evenin Equivalents: For the circuit in Fig. m3.6: (a) Find the open circuit voltage between nodes 1 and 2.

Figure m3.7 Circuit for Problem m3.7.

“book” — 2015/5/4 — 7:08 — page 182 — #68

182 m3.8 Th´evenin Equivalent with Current Source: Creating an ideal current source with the myDAQ requires a current regulator. For the myDAQ portion of this problem, use the LM371 and a 1 k resistor to create the current source in Fig. m3.8. (a) Determine the open circuit voltage. (b) Determine the short circuit current between the output terminals. (c) Determine the Th´evenin resistance for the circuit.

CHAPTER 3 ANALYSIS TECHNIQUES

I1

1.25 mA

R2 1 kΩ

R1

R3

1

1 kΩ

+

2.2 kΩ

Voc

_ 2 Figure m3.8 Circuit for Problem m3.8.

“book” — 2015/5/4 — 7:10 — page 183 — #1

4

CHAPTER

Operational Amplifiers Contents 4-1 TB9 4-2 4-3 4-4 4-5 TB10 4-6 4-7 4-8 4-9 4-10 4-11 TB11 4-12 4-13

Overview, 184 Op-Amp Characteristics, 184 Display Technologies, 190 Negative Feedback, 195 Ideal Op-Amp Model, 196 Inverting Amplifier, 198 Inverting Summing Amplifier, 200 Computer Memory Circuits, 203 Difference Amplifier, 206 Voltage Follower/Buffer, 208 Op-Amp Signal-Processing Circuits, 209 Instrumentation Amplifier, 214 Digital-to-Analog Converters (DAC), 216 The MOSFET as a Voltage-Controlled Current Source, 219 Circuit Simulation Software, 225 Application Note: Neural Probes, 229 Multisim Analysis, 230 Summary, 235 Problems, 236

Dot next to pin #1

4

3

2

1

774411 N SSN 5

6

7

8

The introduction of the operational amplifier chip in the 1960s has led to the development of a wide array of signal processing circuits, enabling the creation of an ever-increasing number of electronic applications.

Objectives Learn to:



Combine multiple op-amp circuits together to perform signal processing operations. Analyze and design high-gain, high-sensitivity instrumentation amplifiers.



Describe the basic properties of an op amp and state the constraints of the ideal op-amp model.





Explain the role of negative feedback and the trade-off between circuit gain and dynamic range.





Analyze and design inverting amplifiers, summing amplifiers, difference amplifiers, and voltage followers.

 

Design an n-bit digital-to-analog converter. Use the MOSFET in analog and digital circuits. Apply Multisim to analyze and simulate circuits that include op amps.

“book” — 2015/5/4 — 7:10 — page 184 — #2

184

CHAPTER 4

Overview Since its first realization by Bob Widlar in 1963 and then its introduction by Fairchild Semiconductor in 1968, the operational amplifier, or op amp for short, has become the workhorse of many signal-processing circuits. It acquired the adjective operational because it is a versatile device capable not only of amplifying a signal but also inverting it (reversing its polarity), integrating it, or differentiating it. When multiple signals are connected to its input, the op amp can perform additional mathematical operations—including addition and subtraction. Consequently, op-amp circuits often are cascaded together in various arrangements to support a variety of different applications. In this chapter, we explore several op-amp circuit configurations, including amplifiers, summers that add multiple signals together, and digital-to-analog converters that convert signals from digital format to analog.

4-1 Op-Amp Characteristics The internal architecture of an op-amp circuit consists of many interconnected transistors, diodes, resistors and capacitors

OPERATIONAL AMPLIFIERS

(Fig. 4-1), all fabricated on a chip of silicon. Despite its internal complexity, however, an op amp can be modeled in terms of a relatively simple equivalent circuit that exhibits a linear inputoutput response. This equivalence allows us to apply the tools we developed in the preceding chapters to analyze (as well as design) a large array of op-amp circuits and to do so with relative ease.

4-1.1

Nomenclature

Commercially available op amps are fabricated in encapsulated packages of various shapes. A typical example is the eightpin DIP configuration shown in Fig. 4-2(a) [DIP stands for dual-in-line package]. The pin diagram for the op amp is shown in Fig. 4-2(b), and its circuit symbol (the triangle) is displayed in Fig. 4-2(c). Of the eight pins (terminals) only five need to be connected to an outside circuit in order for the op amp to function (the remaining three are used for specialized applications). The op amp has two input voltage terminals (υp and υn ) and one output voltage terminal (υo ).

Figure 4-1: The circuit diagram of the Model 741 op amp consists of 20 transistors, several resistors, and one capacitor.

“book” — 2015/5/4 — 7:10 — page 185 — #3

4-1

OP-AMP CHARACTERISTICS

185 to the op amp, KCL mandates that

Op-Amp Pin Designation Pin 2 Pin 3 Pin 4 Pin 7 Pin 6

inverting (or negative) input voltage, υn noninverting (or positive) input voltage, υp negative (−) terminal of power supply Vcc positive (+) terminal of power supply Vcc output voltage, υo

 The terms used to describe pins 3 and 2 as noninverting and inverting are associated with the property of the op amp that its output voltage υo is directly proportional to both the noninverting input voltage υp and the negative of the inverting input voltage υn .  Kirchhoff’s current law applies to any volume of space, including an op amp. Hence, for the five terminals connected

io = ip + in + i+ + i− ,

(4.1)

where ip , in , and io may be constant (dc) or time-varying currents. Currents i+ and i− are dc currents generated by the dc power supply Vcc .  From here on forward, we will ignore the pins connected to Vcc when we draw circuit diagrams involving op amps, because so long as the op amp is operated in its linear region, Vcc will have no bearing on the operation of the circuit.  Hence, the op-amp triangle often is drawn with only three terminals, as shown in Fig. 4-2(d). Moreover, voltages υp , υn , and υo are defined relative to a common reference or ground.

Dot next to pin #1

4

3

2

1

774411 N SSN

1

5 (a)

7 + Vcc (power supply)

υn 2 υp 3 7

6

8

8

+

6 υo

(power supply) −Vcc 4

Typical op-amp package

(b)

5 Pin diagram

i+ ip

υp

in

3 2

+ _

6

io

ip

4

υn

+ (c)

7

i− Vcc

+

Complete circuit diagram

υo

υp υn

Vcc

(d)

+ _

io

in

Op-amp diagram without showing Vcc sources explicitly

Figure 4-2: Operational amplifier.

υo

“book” — 2015/5/4 — 7:10 — page 186 — #4

186

CHAPTER 4

Rs

Positive saturation region (+ voltage rail)

υo Vcc

υs

Maximum negative threshold

+ _

OPERATIONAL AMPLIFIERS

Op-amp circuit with gain G

+ RL

υL = Gυs _

0

Negative saturation region (– voltage rail)

Maximum positive threshold

υp − υn

to the signal input voltage υs .

Linear range Figure 4-3: Op-amp transfer characteristics. The linear range extends between υo = −Vcc and +Vcc . The slope of the line is the op-amp gain A

The (+) and (−) labels printed on the op-amp triangle simply denote the noninverting and inverting pins of the op amp not the polarities of υp or υn . Ignoring the pins associated with the power-supply voltage Vcc does not mean we can ignore currents i+ and i− . To avoid making the mistake of writing a KCL equation on the basis of the simplified diagram given in Fig. 4-2(d), we explicitly state that fact by writing (4.2)

4-1.2 Transfer Characteristics The output voltage υo of the op amp depends on the difference (υp − υn ) at the input side. The plot shown in Fig. 4-3, which depicts the input-output voltage-transfer characteristic of the op amp, is divided into three regions of operation, denoted the negative saturation, linear, and positive saturation regions. In the linear region, the output voltage υo is related to the input voltages υp and υn by υo = A(υp − υn ),

Output load

Figure 4-4: Circuit gain G is the ratio of the output voltage υL

−Vcc

io  = ip + in .

Input circuit

(4.3)

where A is called the op-amp gain, or the open-loop gain. The output voltage can be either positive or negative depending on whether υp is larger than υn or the other way around. Strictly

speaking, this relationship is valid only when the op amp is not connected to an external circuit on the output side (open loop), but as will become clearer in future sections, it continues to hold (approximately) if the output circuit satisfies certain conditions (has high enough input resistance so as not to load the circuit). The open-loop gain is specific to the op-amp device itself, in contrast with the circuit gain or closed-loop gain G, which defines the gain of the entire circuit. Thus, if υs is the signal voltage of the circuit connected at the input side of the op-amp circuit (Fig. 4-4), and υL is the voltage across the load connected at its output side, then υL = Gυs .

(4.4)

According to Eq. (4.3), υo is related linearly to the difference between υp and υn or to either one of them if the other is held constant. Excluding circuits that contain magnetically coupled transformers, in a regular circuit no voltage can exceed the net voltage level of the power supply.  The maximum value that υo can attain is |Vcc |. The op amp goes into a saturation mode if |A(υp − υn )| > |Vcc |, which can occur on both the negative and positive sides of the linear region.  As we will discuss shortly, the op-amp gain A is typically on the order of 105 or greater, and the supply voltage is on the order of volts or tens of volts. In the linear region, υo is bounded between −Vcc and +Vcc , which means that (υp −υn ) is bounded between −Vcc /A and +Vcc /A. For Vcc = 10 V and A = 106 , the operating range of (υp –υn ) is −10 μV to +10 μV. So a basic op-amp configuration is able to amplify only very small voltages, but the configuration can be modified so as to amplify a wider range of voltages (Section 4-2). Even in such cases, however, the maximum output voltage is Vcc and the minimum

“book” — 2015/5/4 — 7:10 — page 187 — #5

4-1

OP-AMP CHARACTERISTICS

187

is −Vcc . These are called the voltage rails. It is important to keep this in mind as we deal with circuits containing operational amplifiers.

4-1.3

Op-Amp Switch

An op amp is an active device. Hence, to operate, it needs to be connected to a power supply that can provide the necessary voltages. Specifically, the op amp requires a positive supply voltage Vcc at pin 7 and a negative supply voltage −Vcc at pin 4. The magnitude of Vcc is specified by the manufacturer. For some models, the positive and negative supply voltages need not be of the same magnitude, but most often they are. Hence, our default assumption in all future considerations of op-amp circuits is that the dc supply voltages connected to pins 4 and 7 are equal in magnitude and opposite in polarity. Among various op-amp models, Vcc typically is between 5 and 24 V. As noted earlier in connection with Fig. 4-4, if (υp − υn ) exceeds a certain maximum positive threshold, the output voltage υo saturates at Vcc , and if (υp − υn ) is negative (because the voltage connected to υp is smaller than that connected to υn ) and its magnitude exceeds a maximum negative threshold, then υo saturates at −Vcc . This op-amp behavior can be used to operate the op amp like an electronic switch, either as an ON/OFF switch, or as a switch to activate one device versus another. An example is illustrated by the circuit in Fig. 4-5. At

Vp Vn

+ _

the input side, the positive terminal is connected to a dc voltage Vp that can be set at either +2 V or −2 V, and the negative input terminal is connected to ground. At the output side, the op amp is connected to the parallel combination of two LEDs, one that can emit red light and another that can emit green light. The two LEDs are arranged in opposite directions, so that when V0 is positive and sufficiently large to cause a current to flow through the red LED, it lights up, but the green LED will neither conduct nor emit green light because it is reverse biased relative to V0 . This is the scenario depicted in Fig. 4-5(b); the input Vp = +2 V (and Vn = 0) causes the output to saturate at V0 = Vcc = 12 V (the vertical flag with Vcc = 12 V is used to denote that this LED uses a Vcc = 12 V), which is quite sufficient to cause the red LED to conduct. When Vp is switched to −2 V, as in the scenario depicted in Fig. 4-5(c), the output saturates at V0 = −12 V, in which case the green LED starts to conduct and emit green light and the red LED stops conducting altogether. Thus, switching the input of the op amp between +2 V and −2 V causes the two LEDs to alternate roles between active and inactive.

4-1.4

Equivalent-Circuit Model in Linear Region

When operated in its linear region, the op-amp input-output behavior can be modeled in terms of the equivalent linear circuit shown in Fig. 4-6. The equivalent circuit consists of a voltagecontrolled voltage source of magnitude A(υp − υn ), an input resistance Ri , and an output resistance Ro . Table 4-1 lists the

Vcc = 12 V

V0

−Vcc = −12 V

R

Red LED

R Green LED

(a) Op-amp circuit

Vp = 2 V

+ _

Vcc = 12 V

−Vcc = −12 V

V0 = 12 V R

Red LED (b) Vp = +2 V

Green LED

Vp = −2 V R LED acts like open circuit

+ _

Vcc = 12 V

V0 = −12 V

−Vcc = −12 V Red LED acts like open circuit

R

R Green LED

(c) Vp = −2 V

Figure 4-5: Op amp operated as a switch. The ±Vcc flags indicate the dc supply voltages connected to pins 7 and 4.

“book” — 2015/5/4 — 7:10 — page 188 — #6

188

CHAPTER 4

OPERATIONAL AMPLIFIERS

Table 4-1: Characteristics and typical ranges of op-amp parameters. The rightmost column represents the values assumed by the ideal op-amp model. Op-Amp Characteristics • Linear input-output response • High input resistance • Low output resistance • Very high gain

Parameter

Typical Range

Ideal Op Amp

Open-loop gain A Input resistance Ri Output resistance Ro Supply voltage Vcc

104 to 108 (V/V) 106 to 1013 

∞ ∞ 0 As specified by manufacturer

typical range of values that each of these op-amp parameters may assume. Based on these values, we note that an op amp is characterized by: (1) High input resistance Ri : at least 1 M, which is highly desirable from the standpoint of voltage transfer from an input circuit (as discussed previously in Section 3-7). (2) Low output resistance Ro : which is desirable from the standpoint of transfering the op-amp’s output voltage to a load circuit. (3) High open loop voltage gain A: which is the key, as we see later, to allowing us to further simplify the equivalent circuit into an “ideal” op-amp model with infinite gain.

1 to 100  5 to 24 V

Example 4-1: Noninverting Amplifier

The circuit shown in Fig. 4-7 uses an op amp to amplify the input signal voltage υs . The circuit uses feedback to connect the op-amp output (at node a) to the inverting input terminal υn through a resistor R1 . Obtain an expression for the circuit gain G = υo /υs , and then evaluate it for Vcc = 10 V, A = 106 , Ri = 10 M, Ro = 10 , R1 = 80 k, and R2 = 20 k. Solution: For reference purposes, we label the output as terminal a and the node from which a current is fed back into the op amp as terminal b. The current i3 flowing from terminal b to terminal a is the same as the current i4 flowing from terminal a towards Ro . (The presence of the voltmeter used to measure υo has no impact on the operation of the circuit because of the very high input resistance of the voltmeter.) When expressed in terms of node voltages, the equality i3 = i4 gives υo − A(υp − υn ) υn − υo = R1 Ro

(node a).

(4.5)

At node b, KCL gives i1 + i2 + i3 = 0, or

υp ip

+

in υn

υn − υp υn υn − υo + + = 0. Ri R2 R1

+ Ro (υp − υn) Ri A(υp − υn) − + _ − −

io

+

(node b).

(4.6)

Additionally,

υo

υp = υs .

(4.7)

Solution of these simultaneous equations leads to the following expression for the circuit gain G: υo [ARi (R1 + R2 ) + R2 Ro ] = . υs AR2 Ri + Ro (R2 + Ri ) + R1 R2 + Ri (R1 + R2 ) (4.8) For Vcc = 10 V, A = 106 , Ri = 107 , Ro = 10 , R1 = 80 k, and R2 = 20 k, G=

Figure 4-6: Equivalent circuit model for an op amp operating in the linear range (υo ≤ |Vcc |). Voltages υp , υn , and υo are referenced to ground.

G=

υo = 4.999975 ≈ 5.0. υs

(4.9)

“book” — 2015/5/4 — 7:10 — page 189 — #7

4-1

OP-AMP CHARACTERISTICS

189

Vcc = 10 V

υp

υs

υo ≈

+

υn

i4

Ro

Ri

+ _

R1 + R2 υs R2

(

)

a

A(υp − υn) + _

_

+ R1 −Vcc = −10 V i1

i3

υo

b

Negative feedback (connecting output to negative input terminal)

υn

i2

_

R2

Figure 4-7: Noninverting amplifier circuit of Example 4-1. In the expression for G, the two parameters A and Ri are several orders of magnitude larger than all of the others. Also, Ro is in series with R1 , which is 8000 times larger. Hence, we would incur minimal error if we let A → ∞, Ri → ∞, and Ro → 0, in which case the expression for G reduces to G≈

R1 + R2 R2

(ideal op-amp model).

(4.10)

This approximation, based on the ideal op-amp model that will be introduced in Section 4-3, gives G=

80 k + 20 k = 5. 20 k

Concept Question 4-3: How is an op amp used as a

switch? (See

)

Concept Question 4-4: An op amp is characterized by three important input-output attributes. What are they? (See )

Exercise 4-1: In the circuit of Example 4-1 shown in Fig. 4-7, insert a series resistance Rs between υs and υp and then repeat the solution to obtain an expression for G. Evaluate G for Rs = 10  and use the same values listed in Example 4-1 for the other quantities. What impact does the insertion of Rs have on the magnitude of G? Answer:

Concept Question 4-1: How is the linear range of an op

amp defined? (See

G=

)

[A(Ri + Rs )(R1 + R2 ) + R2 Ro ] [AR2 (Ri + Rs ) + Ro (R2 + Ri + Rs ) + R1 R2 + (Ri + Rs )(R1 + R2 )]

= 4.999977 Concept Question 4-2: What is the difference between

the op-amp gain A and the circuit gain G? (See

)

(See

)

(negligible impact).

“book” — 2015/5/4 — 7:10 — page 190 — #8

190

TECHNOLOGY BRIEF 9: DISPLAY TECHNOLOGIES

Technology Brief 9 Display Technologies

• During operation, the cathode emits streams of electrons into the electron gun.

From cuneiform-marked clay balls to the abacus to today’s digital projection technology, advances in visual displays have accompanied almost every major leap in information technology. While the earliest “modern” computers relied on cathode ray tubes (CRT) to project interactive images, today’s computers can access a wide variety of displays ranging from plasma screens and LED arrays to digital micromirror projectors, electronic ink, and virtual reality interfaces. In this Technology Brief, we will review the major technologies currently available for twodimensional visual displays.

• The emitted electron stream is steered onto different parts of the positively charged screen by the electron gun; the direction of the electron stream is controlled by the electric field of the deflecting coils through which the beam passes. • The screen is composed of thousands of tiny dots of phosphorescent material arranged in a twodimensional array. Every time an electron hits a phosphor dot, it glows a specific color (red, blue, or green). A pixel on the screen is composed of phosphors of these three colors.

Cathode Ray Tube (CRT)

• In order to make an image appear to move on the screen, the electron gun constantly steers the electron stream onto different phosphors, lighting them up faster than the eye can detect the changes, and thus, the images appear to move. In modern color CRT displays, three electron guns shoot different electron streams for the three colors.

The earliest computers relied on the same technology that made the television possible. In a CRT television or monitor (Fig. TF9-1), an electron gun is placed behind a positively charged glass screen, and a negatively charged electrode (the cathode) is mounted at the input of the electron gun.

Electron beam

Deflecting coil Anodes

Electron-emitting heated cathode

Light emitted from phosphor Focusing anode Deflecting coil

Evacuated glass enclosure Figure TF9-1: Schematic of CRT operation.

“book” — 2015/5/4 — 7:10 — page 191 — #9

TECHNOLOGY BRIEF 9: DISPLAY TECHNOLOGIES

Horizontal polarization filter

191

Glass

Front display glass with color filter

Polarized light

Row and column electrodes

Vertical polarization filter

Figure TF9-2: Schematic of LCD operation.

The basic concept behind CRT was explored in the early 2000s in the development of field emission displays (FED), which used a thin film of atomically sharp electron emitter tips to generate electrons. The electrons emitted by the film collide with phosphor elements just as in the traditional CRT. The primary advantage of this type of “flat-panel” display is that it can provide a wider viewing angle (i.e., one can look at an FED screen at a sharp angle and still see a good image) than possible with conventional LCD or LED technology (discussed next).

Liquid Crystal Displays (LCD) LCDs are used in digital clocks, cellular phones, desktop and laptop computers, and some televisions and other electronic systems. They offer a decided advantage over other display technologies (such as cathode ray tubes) in that they are lighter and thinner and consume a lot less power to operate. LCD technology relies on special electrical and optical properties of a class of materials known as liquid crystals, first discovered in the 1880s by botanist Friedrich Reinitzer. In the basic LCD display, light shines through a thin stack of layers as shown in Fig. TF9-2. • Each stack consists of layers in the following order (starting from the viewer’s eye): color filter, vertical (or horizontal) polarizer filter, glass plate with transparent electrodes, liquid crystal layer, second

glass plate with transparent electrodes, horizontal (or vertical) polarizer filter. • Light is shone from behind the stack (called the backlight). As light crosses through the layer stack, it is polarized along one direction by the first filter. • If no voltage is applied on any of the electrodes, the liquid crystal molecules align the filtered light so that it can pass through the second filter. • Once through the second filter, it crosses the color filter (which allows only one color of light through) and the viewer sees light of that color. • If a voltage is applied between the electrodes on the glass plates (which are on either side of the liquid crystal), the induced electric field causes the liquid crystal molecules to rotate. Once rotated, the crystals no longer align the light coming through the first filter so that it can pass through the second filter plate. • If light cannot cross, the area with the applied voltage looks dark. This is precisely how simple hand-held calculator displays work; usually the bright background is made dark every time a character is displayed. Modern monitors, laptops, phones, and tablets use a version of the LCD called thin-film transistor (TFT) LCD; these also are known as active matrix displays. In TFT

“book” — 2015/5/4 — 7:10 — page 192 — #10

192 LCDs, several thin films are deposited on one of the glass substrates and patterned into transistors. Each color component of a pixel has its own microscale transistor that controls the voltage across the liquid crystal; since the transistors only take up a tiny portion of the pixel area, they effectively are invisible. Thus, each pixel has its own electrode driver built directly into it. This specific feature enabled the construction of the flat high-resolution screens now in common use (and made the CRT display increasingly obsolete). Since LCD displays also weigh considerably less than a CRT tube, they enabled the emergence of laptop computers in the 1980s. Early laptops used large, heavy monochrome LCDs; most of today’s mobile devices use active-matrix displays.

Light-Emitting Diode (LED) Displays A different but very popular display technology employs tiny light-emitting diodes (LED) in large pixel arrays on flat screens (see Technology Brief 5 on LEDs). Each pixel in an LED display is composed of three LEDs (one each of red, green, and blue). Whenever a current is made to pass through a particular LED, it emits light at its particular color. In this way, displays can be made flatter (i.e., the LED circuitry takes up less room than an electron gun or LCD) and larger (since making large, flat LED arrays technically is less challenging than giant CRT tubes or LCD displays). Unlike LCDs, LED displays do not need a backlight to function and easily can be made multicolor. Modern LED research is focused mostly on flexible and organic LEDs (OLEDs), which are made from polymer light-emitting materials and can be fabricated on flexible substrates (such as an overhead transparency). Flexible displays of this type have been demonstrated by several groups around the world.

Plasma Displays Plasma displays have been around since 1964 when invented at the University of Illinois.While attractive due to their low profile, large viewing angle, brightness, and large screen size, they largely were displaced in the 1980s in the consumer market by LCD displays for manufacturingcost reasons. In the late 1990s, plasma displays became popular for high-definition television (HDTV) systems. Each pixel in a plasma display contains one or more microscale pocket(s) of trapped noble gas (usually neon or xenon); electrodes patterned on a glass substrate are placed in front and behind each pocket of gas (Fig.TF9-3).

TECHNOLOGY BRIEF 9: DISPLAY TECHNOLOGIES

Plasma cells with phosphors

Insulator

Light

Row and column electrodes

Front display glass

Figure TF9-3: Plasma display.

The back of one of the glass plates is coated with light-emitting phosphors. When a sufficient voltage is applied across the electrodes, a large electric field is generated across the noble gas, and a plasma (ionized gas) is ignited. The plasma emits ultraviolet light which impacts the phosphors; when impacted with UV light, the phosphors emit light of a certain color (blue, green, or red). In this way, each pocket can generate one color.

Electronic Ink Electronic ink, e-paper, or e-ink are all names for a set of display technologies made to look like paper with ink on it. In all cases, the display is very thin (almost as thin as real paper), does not use a backlight (ambient light is reflected off the display, just like real paper), and little to no power is consumed when the image is kept constant. The first version of e-paper was invented in the 1970s at Xerox, but it was not until the 1990s that a commercially viable version was developed at MIT. A number of electronic ink technologies are in production or in development. • Most common electronic ink technologies trap a thin layer of oil between two layers of glass or plastic onto which have been patterned transparent electrodes. The total stack is usually less than a tenth of a millimeter. • Within the oil are suspended charged particles. In some versions, the oil is colored.

“book” — 2015/5/4 — 7:10 — page 193 — #11

TECHNOLOGY BRIEF 9: DISPLAY TECHNOLOGIES

193

Table TT9-1: A comparison of some characteristics of common display technologies; see also http://en.wikipedia.org/wiki/ Comparison of CRT, LCD, Plasma, and OLED.

Pros • Good dynamic range (~15,000 : 1) • Very little distortion • Excellent viewing angle • No inherent pixels

Cons Cathode Ray Tube (CRT) • Large and heavy, limiting maximum practical size • High power consumption and heat generation • Burn-in possible • Produces noticeable flicker at low refresh rates • Minimum size for color limited to 7” diagonal • Can contain lead, barium, and cadmium, which are toxic

Plasma Displays • Excellent contrast ratios (~1,000,000 : 1) • Large minimum pixel pitch; suitable for larger displays • Sub-millisecond response time • High power consumption than LCD • Near zero distortion • Limited color depth since plasma pixels can only be turned on or off, no grading of emission • Excellent viewing angle • Very scalable (easier than other technologies to make large • Image burn-in possible displays)

Organic Light-Emitting Diode (LED) Displays • Excellent viewing angle • Limited lifetime of organic materials (but progress in this area is rapid) • Very light • Very fast, so no image distortion during fast motion • Burn-in possible • Excellent color quality because no backlight is used • More expensive than other technologies (ca. 2012) Liquid Crystal Displays (LCD) • Small and light • Limited viewing angle • Lower power consumption than plasma or CRT • Slower response than plasma or CRT can cause image distortion during fast motion • No geometric distortion • Can be made in almost any size or shape • Slow response at low temperatures • Liquid crystal has no inherent resolution limit • Requires a backlight, which can vary across screen Digital Light Projection (DLP) Displays • No burn-in • Requires light source replacement • Cheaper than LCD or plasma displays • Reduced viewing angle compared with CRT, plasma, and LCD • DLPs with LED and laser sources do not need light source • Some viewers perceive the colors in the projection, replacement very often

producing a rainbow effect

• Excellent for very large screens (theaters) due to possibility of using multiple color sources (color depth) and no inherent size limitation to hardware

Electronic Ink Displays • Very low power consumption • Slow, consumer units not yet suitable for fast video • Works with reflected light; excellent for viewing in bright light • Ghost images persist without refresh • Lightweight • Color displays are still under development • Flexible and bendable • Applying a potential across the electrodes on either side of the oil suspension attracts the charged particles to either the top or bottom substrates (depending on the polarity). Some displays use white

particles in black fluid. Thus, when the white particles move to the top, they block the black fluid and the display appears white. When they move to the bottom, the display appears dark. Some displays use

“book” — 2015/5/4 — 7:10 — page 194 — #12

194

TECHNOLOGY BRIEF 9: DISPLAY TECHNOLOGIES

Micromirror pixel Digital micromirror chip Lens

Projected light

Lens Light source

Figure TF9-4: A typical digital light processor (DLP) arrangement includes a light source, lenses, and a micromirror array that steers the light to create projected pixels.

a combination of black and white particles to achieve the same effect.

Digital Light Processing (DLP) Digital light processing (DLP) is the name given to a technology that uses arrays of individual, micromechanical mirrors to manipulate light at each pixel position. Invented in 1987 by Dr. Hornbeck at Texas Instruments, this technology has revolutionized projection technology; many of today’s digital projectors are made possible by DLP chips. DLP also was used heavily in large, rear-projection televisions. • A basic DLP consists of an array of metal micromirrors, each about 100 micrometers on a side (Fig. TF9-4(inset)). One micromirror corresponds to one pixel on a digital image. • Each micromirror is mounted on micromechanical hinges and can be tilted towards or away from a light source several thousand times per second!

• The mirrors are used to reflect light from a light source (housed within the television or projector case) and through a lens to project it either from behind a screen (as is the case in rear-projection televisions) or onto a flat surface (in the case of projectors), as in (Fig. TF9-4). If a micromirror is tilted away from the light source, that pixel on the projected image becomes dark (since the mirror is not passing the light onto the lens).

• If it is tilted towards the light source, the pixel lights up. By varying the relative time a given mirror is in each position, grey values can be generated as well.

• Color can be added by using multiple light sources and either one chip (with a filter wheel) or three chips. The three-chip color DLP used in high-resolution cinema systems can purportedly generate 35 trillion different colors!

“book” — 2015/5/4 — 7:10 — page 195 — #13

4-2

NEGATIVE FEEDBACK

195

4-2 Negative Feedback

with G≈

 Feedback refers to taking a part of the output signal and feeding it back into the input. It is called positive feedback if it increases the intensity of the input signal, and it is called negative feedback if it decreases it. In negative feedback, the output terminal is connected to the υn terminal, either directly or through a resistor. 

Positive feedback causes the op amp to saturate, thereby forcing its output voltage υo to become equal to its supply voltage Vcc . This behavior is used to advantage in certain types of applications but they are outside the scope of this book. Negative feedback, on the other hand, is an essential ingredient of all of the op-amp circuits covered in this and forthcoming chapters. Why do some op-amp circuits need feedback and why negative feedback specifically? It seems counter-intuitive to want to decrease the input signal when the intent is to amplify it! We will answer this question by examining the circuit of Example 4-1 in some detail. To facilitate the discussion we have reproduced the circuit diagram (into a smaller version) and inserted it in Fig. 4-8(a). When we say an op amp has a supply voltage Vcc of 10 V, we actually mean that a positive (10 V) dc voltage source is connected to pin 7 of its package and another, negative (−10 V) source is connected to its pin 4 (Fig. 4-2(b)). The op-amp circuit cannot generate an output voltage υo that exceeds its supply voltage. Hence, υo is bounded to ±Vcc which means

|υo | ≤ Vcc ,

R1 + R2 . R2

(4.13)

Inserting Eq. (4.12) into Eq. (4.11) gives |Gυs | ≤ Vcc ,

(4.14)

−Vcc Vcc ≤ υs ≤ , G G

(4.15)

or

which states that the linear dynamic range of υs is inversely proportional to the circuit gain G. (a) Unity Gain: If R2 = ∞ (open circuit between node b and ground in the circuit of Fig. 4-8(a)), Eq. (4.13) gives G ≈ 1. The corresponding dynamic range of υs extends from −Vcc to +Vcc , the same as the output. The input-output transfer plot relating υo to υs is displayed in green in Fig. 4-8(b). (b) Modest Gain: If we choose R1 /R2 = 4, Eq. (4.13) gives G = 5, and the dynamic range of υs now extends from −(10/5) = −2V to +2 V. Thus, the gain is higher than the unity-gain case by a factor of 5, but the dynamic range of υs is narrower by the same factor. (c) Maximum Gain: If R1 is removed (replaced with an open circuit between nodes a and b) and R2 is set equal to zero (short circuit), no feedback will take place in the circuit of Fig. 4-8(a). Use of the exact expression for G given by Eq. (4.8) leads to G = A. Since A = 106 , the absence of feedback provides a huge gain, but operationally υs becomes limited to a very narrow range extending from −10 μV to +10 μV.  Application of negative feedback offers a trade-off between circuit gain and dynamic range for the input voltage. 

or equivalently, −Vcc ≤ υo ≤ Vcc .

(4.11) Concept Question 4-5: Why is negative feedback used

Thus, the linear dynamic range of υo extends from −Vcc to +Vcc . According to Example 4-1, υo is related to the signal voltage υs by υo = Gυs ,

(4.12)

in op-amp circuits? (See

)

Concept Question 4-6: How large is the circuit gain G in the absence of feedback? How large is it with 100 percent feedback (equivalent to setting R1 = 0 in the circuit of Fig. 4-8(a))? (See )

“book” — 2015/5/4 — 7:10 — page 196 — #14

196

CHAPTER 4

OPERATIONAL AMPLIFIERS

υ0 Vcc

Vcc = 10 V υp

υs

+ _

+ i4 a

Ro

Ri υn

G = 1, |υs| < Vcc G = 5, |υs| < Vcc / 5 G = 10, |υs| < Vcc / 10

+ _

_

+

A(υp − υn)

υs

υo

R1 −Vcc i1 Feedback

i3 b

υn

i2

Dynamic range (high gain)

R2

Dynamic range (modest gain) Dynamic range (unity gain)

(a)

(b) Input-output transfer plots Figure 4-8: Trade-off between gain and dynamic range.

Exercise 4-2: To evaluate the trade-off between the

circuit gain G and the linear dynamic range of υs , apply Eq. (4.8) to find the magnitude of G and then determine the corresponding dynamic range of υs for each of the following values of R2 : 0 (no feedback), 800 , 8.8 k, 40 k, 80 k, and 1 M. Except for R2 , all other quantities remain unchanged. R2

Answer:

0 800  8.8 k 40 k 80 k 1 M

(See

)

G

υs Range

106 101 10.1 3 2 1.08

−10 μV to +10 μV −99 mV to +99 mV −0.99 V to +0.99 V −3.3 V to +3.3 V −5 V to +5 V −9.26 V to +9.26 V

4-3

Ideal Op-Amp Model

We noted in Section 4-1 that the op amp has a very large input resistance Ri on the order of 107 , a relatively small output resistance Ro on the order of 1–100 , and an openloop gain A ≈ 106 . Usually, the series resistances of the input circuit connected to terminals υp and υn are several orders of magnitude smaller than Ri . Consequently, not only will very little current flow through the input circuit, but also the voltage drop across the input-circuit resistors will be negligibly small in comparison with the voltage drop across Ri . These considerations allow us to simplify the equivalent circuit of the op amp by replacing it with the ideal op-amp circuit model shown in Fig. 4-9, in which Ri has been replaced with an open circuit. An open circuit between terminals υp and υn implies

“book” — 2015/5/4 — 7:10 — page 197 — #15

4-3

IDEAL OP-AMP MODEL

197

Table 4-2: Characteristics of the ideal op-amp model.

ip = 0

+ in = 0

υn

(Ri =

_

)

8

υp

(Ro = 0)

+

Ideal Op Amp

υo

• Current constraint • Voltage constraint • A = ∞ Ri = ∞

ip = in = 0 υp = υn Ro = 0

Noninverting Amplifier Figure 4-9: Ideal op-amp model.

Rs

υn

the following ideal op-amp current constraint: ip = in = 0

(ideal op-amp model).

(4.16)

In reality, ip and in are very small but not identically zero; for if they were, there would be no amplification through the op amp. Nevertheless, the current condition given by Eq. (4.16) will prove quite useful. Similarly, at the output side, if the load resistor connected in series with Ro is several orders of magnitude larger than Ro , then Ro can be ignored by setting it equal to zero. Finally, in the ideal op-amp model, the large open-loop gain A is made infinite—the consequence of which is that υo υp − υn = →0 A

υp ip = 0

as A → ∞.

υs

+ _

+

υo

in = 0 −

R1

Rinput

υn R2

υp = υn Rinput ≈ Ri ≈ ∞ (a)

Circuit

υs (b)

G=

R1 + R2 R2

υo = Gυs

Block-diagram representation

Figure 4-10: Noninverting amplifier circuit: (a) using ideal op-amp model and (b) equivalent block-diagram representation.

Hence, we obtain the ideal op-amp voltage constraint υp = υn

(ideal op-amp model).

(4.17)

In reality υp and υn are not exactly equal, but very close to being equal, and only when negative feedback is in use. Nevertheless, setting υp = υn leads to highly accurate results when relating the output to the input. In summary:

To illustrate the utility of the ideal op-amp model, let us reexamine the circuit we analyzed earlier in Example 4-1, but we will do so this time using the ideal model. The new circuit, as shown in Fig. 4-10, includes a source resistance Rs , but because the op amp draws no current (ip = 0), there is no voltage drop across Rs . Hence, υp = υs ,

(4.18)

 The ideal op-amp model characterizes the op amp in terms of an equivalent circuit in which Ri = ∞, Ro = 0, and A = ∞. 

and on the output side, υo and υn are related through voltage division by   R1 + R2 υn . (4.19) υo = R2

The operative consequences are given by Eqs. (4.16) and (4.17) and in Table 4-2.

Using these two equations, in conjunction with υp = υn (from Eq. (4.17)), we end up with the following result for the circuit

“book” — 2015/5/4 — 7:10 — page 198 — #16

198

CHAPTER 4

4-4

gain G: G=

υo = υs



R1 + R2 R2

OPERATIONAL AMPLIFIERS

Inverting Amplifier

 ,

(4.20)

 In an inverting amplifier op-amp circuit, the input source is connected to terminal υn (instead of to terminal υp ) through an input source resistance Rs , and terminal υp is connected to ground. 

which is identical to Eq. (4.10).  The input resistance of the noninverting amplifier circuit shown in Fig. 4-10 is the Th´evenin resistance of the op-amp circuit as seen by the input source υs . Because ip = 0, it is easy to show that Rinput = Ri ≈ ∞, where Ri is the input resistance of the op amp (typically on the order of 109 ).   From here on forward, we use the ideal op-amp model exclusively. 

Feedback from the output continues to be applied at υn (through a feedback resistance Rf ), as shown in Fig. 4-11. It is called an inverting amplifier because (as we will see shortly) the circuit gain G is negative. To relate the output voltage υo to the input signal voltage υs , we start by writing down the node-voltage equation at terminal υn as i1 + i2 + in = 0

(4.21)

υn − υo υn − υs + + in = 0. Rs Rf

(4.22)

or Concept Question 4-7: What are the current and voltage

constraints of the ideal op amp? (See

)

Upon invoking the op-amp current constraint given by Eq. (4.16), namely in = 0, and the voltage constraint υn = υp ,

Concept Question 4-8: What are the values of the input

and output resistances of the ideal op amp? (See

)

Inverting Amplifier Concept Question 4-9: In the ideal op-amp model, Ro

Rf

is set equal to zero. To satisfy such an approximation, does the load resistance need to be much larger or much smaller than Ro? Explain. (See )

i2 in = 0

Exercise 4-3: Consider the noninverting amplifier circuit

Rs

υs

of Fig. 4-10(a) under the conditions of the ideal op-amp model. Assume Vcc = 10 V. Determine the value of G and the corresponding dynamic range of υs for each of the following values of R1 /R2 : 0, 1, 9, 99, 103 , 106 . Answer:

R1 /R2

G

0 1 9 99 1000 106

1 2 10 100 ∼ 1000 ∼ 106

(See

)

υs Range −10 V to +10 V −5 V to +5 V −1 V to +1 V −0.1 V to +0.1 V −10 mV to +10 mV (approx.) −10 μV to +10 μV (approx.)

+ + -_

i1

υn υp ip = 0

Rinput

Feedback −

υo

+

RL

υυp = υυn Rinput ≈ Rs (a)

Circuit

υs (b)

G = − (Rf /Rs)

υo = Gυs

Block diagram

Figure 4-11: Inverting amplifier circuit and its block-diagram equivalent.

“book” — 2015/5/4 — 7:10 — page 199 — #17

4-4

INVERTING AMPLIFIER

199

as well as recognizing that υp = 0 (because terminal υp is connected to ground), we obtain the relationship  υo = −

Rf Rs

Rf

 υs .

The circuit voltage gain of the inverting amplifier therefore is given by   υo Rf G= =− . υs Rs

R1

(4.23)

υn

+

υp

R2

is

+



υo

RL

(4.24)

(a)

Original circuit

 In addition to amplifying υs by the ratio (Rf /Rs ), the inverting amplifier also reverses the polarity of υs . 

Rf Rs = R1 + R2

 υo is independent of the magnitude of the load resistance RL , so long as RL is much larger than the opamp output resistance Ro (which is an implicit assumption of the ideal op-amp model).  Because υn = 0, a Th´evenin analysis of the circuit in Fig. 4-11(a) would reveal that the input resistance of the inverting amplifier circuit (as seen by source υs ) is Rinput = RTh = Rs .

+ _

+



+

υp

υs = isR2

(b)

υn

υo

RL

After source transformation

Figure 4-12: Inverting amplifier circuit of Example 4-2.  Caution: Under the ideal op-amp model, it is not possible to compute io , the current that flows into the op amp from output terminal υo . Hence, it is inappropriate to apply KCL at that terminal because additional current can be delivered by the supply voltage sources Vcc and −Vcc . 

Example 4-2: Amplifier with Input Current Source

For the circuit shown in Fig. 4-12(a): (a) obtain an expression for the input-output transfer function Kt = υo / is and evaluate it for R1 = 1 k, R2 = 2 k, Rf = 30 k, and RL = 10 k; and (b) determine the linear dynamic range of is if Vcc = 20 V. Solution: (a) Application of the source transformation method converts the combination of is and R2 into a voltage source υs = is R2 , in series with a resistance R2 . Upon combining R2 in series with R1 , we obtain the new circuit shown in Fig. 4-12(b), which is identical in form to the inverting amplifier circuit of Fig. 4-11, except that now the source

resistance is Rs = (R1 + R2 ). Hence, application of Eq. (4.23) gives  υo = −

Rf R1 + R 2



 υs = −

Rf R1 + R 2

 R2 is ,

(4.25)

from which we obtain the transfer function Kt =

υo R f R2 =− . is R1 + R 2

For R1 = 1 k, R2 = 2 k, and Rf = 30 k, Kt =

υo = −2 × 104 is

(b) From the expression for Kt , is = −

υo , 2 × 104

(V/A).

(4.26)

“book” — 2015/5/4 — 7:10 — page 200 — #18

200

CHAPTER 4

and since |υo | is bounded by Vcc = 20 V, the linear range for is is bounded by   |is | = 

OPERATIONAL AMPLIFIERS

Inverting Summing Amplifier

   Vcc   20  = = 1 mA. 2 × 104   2 × 104 

R1

Summing point R2

Thus, the linear range of is extends from −1 mA to +1 mA.

υ1

+ _

υ2

+ _

υp

υn

Rf

_ +

υo

Concept Question 4-10: How does feedback control the

gain of the inverting-amplifier circuit? (See

)

Original circuit (a)

Concept Question 4-11: The expression given by

Eq. (4.24) states that the gain of the inverting amplifier is independent of the magnitude of RL. Would the expression remain valid if RL = 0? Explain. (See )

υ1 is1 = R1

Exercise 4-4: The input to an inverting-amplifier circuit

is1

Rf

υ2 is2 = R2 R1

is2

R2

υp

υn

_ +

consists of υs = 0.2 V and Rs = 10 . If Vcc = 12 V, what is the maximum value that Rf can assume before saturating the op amp? Answer: Gmax = −60, Rf = 600 . (See

C3

Rf υ1R2 + υ2R1 υs = R + R 1 2

By connecting multiple sources in parallel at terminal υn of the inverting amplifier, the circuit becomes an adder (or more precisely a scaled inverting adder), as depicted by the block diagram of Fig. 4-13(d). After we demonstrate how such a circuit (usually called an inverting summing amplifier) works for two input voltages υ1 and υ2 , we will extend it to multiple sources. There are many applications where we may want to scale and add multiple voltages together, such as combining or averaging results from several sensors. For the circuit shown in Fig. 4-13(a), our goal is to relate the output voltage υo to υ1 and υ2 . To do so, we apply the source-transformation technique so as to cast the input circuit in the form of a single voltage source υs in series with a source resistance Rs . The steps involved in the transformation are illustrated in Fig. 4-13(b) and (c). Voltage to current transformation gives is1 = υ1 /R1 and is2 = υ2 /R2 , which can be combined together into a single current source as υ1 υ2 υ1 R2 + υ2 R1 + = . R1 R2 R1 R2

After source transformation (b)

)

4-5 Inverting Summing Amplifier

is = is1 + is2 =

υo

(4.27)

υs

Rs

+ _

υp

υn

_ +

(c)

After combining and retransforming

υ1

G1 = − Rf /R1

υ2

G2 = − Rf /R2

(d)

+

υo

υo = G1υ1 + G2υ2

Block diagram representation Figure 4-13: Inverting summing amplifier.

Similarly, the two parallel resistors add up to Rs =

R 1 R2 . R1 + R 2

(4.28)

“book” — 2015/5/4 — 7:10 — page 201 — #19

4-5

INVERTING SUMMING AMPLIFIER

201

If we transform (is , Rs ) into a voltage source (υs , Rs ), we get  υs = is Rs =

υ1 R2 + υ2 R1 R1 R2



R1 R2 υ1 R2 + υ2 R1 = . R1 + R 2 R1 + R 2 (4.29)

The circuit in Fig. 4-13(c) is identical in form to that of the inverting amplifier of Fig. 4-11. Hence, by applying the inputoutput voltage relationship given by Eq. (4.23), we have 

Rf υo = − Rs



  υ1 R2 + υ2 R1 Rf  υs = −  R 1 R2 R1 + R 2 R1 + R 2     Rf Rf =− υ1 − υ2 . (4.30) R1 R2

This expression for υo can be written in the form υo = G1 υ1 + G2 υ2 ,

(4.31)

where G1 = −(Rf /R1 ) is the (negative) gain applied to source voltage υ1 , and G2 = −(Rf /R2 ) is the gain applied to υ2 . Thus:  The summing amplifier scales υ1 by negative gain G1 and υ2 by negative gain G2 and adds them together. 

4-5.1

Generalizing to the case where the input consists of n input voltage sources υ1 to υn (and associated source resistances R1 to Rn , respectively), all connected in parallel at the same summing point (terminal υn ), the output voltage becomes       Rf Rf Rf υ1 + − υ2 + · · · + − υn . υo = − R1 R2 Rn (4.34)

Example 4-3: Summing Circuit

Use inverting amplifiers to design a circuit that performs the operation υo = 4υ1 + 7υ2 . Solution: The desired circuit has to amplify υ1 by a factor of 4, amplify υ2 by a factor of 7, and add the two together. A summing amplifier can do that, but it also inverts the sum. Hence, we will need to use a two-stage cascaded circuit with the first stage providing the desired operation within a “−” sign and then follow it up with an inverting amplifier with a gain of (−1). The two-stage circuit is shown in Fig. 4-14. For the first stage, we need to select values for R1 , R2 , and Rf1 such that Rf1 Rf1 =4 and = 7. R1 R2 Since we have only two constraints, we can satisfy the specified ratios with an infinite number of combinations. Arbitrarily, we choose Rf1 = 56 k, which then specifies the other resistors as

Special Cases

For the special case where R1 = R2 = R,

υo = −



Rf R



 (υ1 + υ2 )

equal gain R1 = R2 = R

R1 = 14 k  ,

(4.32)



 inverted adder . R1 = R2 = Rf

R2 = 8 k.

For the second stage, a gain of (−1) requires that Rf2 = 1. Rs2

and if additionally Rf = R1 = R2 , then G1 = G2 = −1. In this case, the summing amplifier becomes an inverted adder with

υo = −(υ1 + υ2 )

and

Arbitrarily, we choose Rf2 = Rs2 = 20 k.

4-5.2

Noninverting Summer

(4.33) To perform the summing operation, the solution offered in Example 4-3 employed two inverting amplifier circuits—one

“book” — 2015/5/4 — 7:10 — page 202 — #20

202

CHAPTER 4

OPERATIONAL AMPLIFIERS

Rf1 R1

υ1

+ _

Rf2

υn1



R2 υ2

+

υp1

+ _

Rs2 υn 2

υo1

υo1 =

− υo2

υp2

+

(− RR ) υ + (− RR ) υ f1 1

1

f1 2

υo2 =

2

Stage 1: Inverting summing amp

(− RR ) υ f2

s2

o1

Stage 2: Inverting amp

(a) Two-stage circuit Circuit Design R1 R2 Rf1 Rs1 Rf2

14 kΩ 8 kΩ 56 kΩ 20 kΩ 20 kΩ

υ1

−4

υ2

+

υo1

−1

υo2

−7 (b) Block diagram

Figure 4-14: Two-stage circuit realization of υo = 4υ1 + 7υ2 .

to perform an inverted sum, and a second one to provide multiplication by (−1). Alternatively, the same result can be achieved by using a single op amp in a noninverting amplifier circuit, as shown in Fig. 4-15. From our analysis in Section 4-3, we established that the output voltage υo of the noninverting amplifier circuit is related to υp by R1 + R2 υo =G= . υp R2

(4.35)

For the circuit in Fig. 4-15, in view of the ideal op-amp constraint that the op amp draws no current (ip = 0), it is a straightforward task to show that (4.36)

Combining Eqs. (4.35) and (4.36) leads to 

Rs2 Rs2 + Rs1



 υ1 +

Rs1 Rs1 + Rs2

GRs2 =4 Rs1 + Rs2 and GRs1 = 7. Rs1 + Rs2 A possible solution that satisfies these two constraints is Rs1 = 7 k, Rs2 = 4 k, and G = 11. Furthermore, the specified value of G can be satisfied by choosing R1 = 50 k and R2 = 5 k.

4-5.3

υ1 Rs2 + υ2 Rs1 . υp = Rs1 + Rs2

υo = G

To realize a coefficient of 4 for υ1 and a coefficient of 7 for υ2 , it is necessary that



 υ2 .

(4.37)

Multiple Ways of Building a System

There are often several different choices for how to implement a linear equation such as υo = 4υ1 + 7υ2 (Example 4-3) with op-amp circuits. Here are a few options: (a) υo = (4υ1 ) + (7υ2 ): Multiply υ1 by 4 (noninverting amplifier with a gain of 4) and υ2 by 7 (noninverting amplifier with a gain of 7), and then add them together (noninverting summer with a gain of 1).

“book” — 2015/5/4 — 7:10 — page 203 — #21

TECHNOLOGY BRIEF 10: COMPUTER MEMORY CIRCUITS

Technology Brief 10 Computer Memory Circuits The storage of information in electronically addressable devices is one of the hallmarks of all modern computer systems. Among these devices are a class of storage media, collectively called solid-state or semiconductor memories, which store information by changing the state of an electronic circuit. The state of the circuit usually has two possibilities (0 or 1) and is termed a bit (see Technology Brief 8). Values in memories are represented by a string of binary bits; a 5-bit sequence [V1V2V3V4V5 ], for example, can be used to represent any integer decimal value between 0 and 31. How do computers store these bits? Many types of technologies have emerged over the last 40 years, so in this Brief, we will highlight some of the principal technologies in use today or under development. It is worth noting that memory devices usually store these values in arrays. For example, a small memory might store sixteen different 16-bit numbers; this memory usually would be referred to as a 16 × 16 block or a 256bit memory. Of course, modern multi-gigabyte computer memories use thousands of much larger blocks to store very large numbers of bits (Fig. TF10-1).

203

Read-Only Memories (ROMs) One of the oldest, still-employed, memory architectures is the read-only memory (ROM). The ROM is so termed because it can only be “written” once, and after that it can only be read. ROMs usually are used to store information that will not need to be changed (such as certain startup information on your computer or a short bit of code always used by an integrated circuit in your camera). Each bit in the ROM is held by a single MOSFET transistor. Consider the circuit in Fig. TF10-2(a), which operates much like the circuit in Fig. 4-25. The MOSFET has three voltages, all referenced to ground. For convenience, the input voltage is labeled VREAD and the output voltage is labeled VBIT . The third voltage, VDD , is the voltage of the dc power supply connected to the drain terminal via a resistor R. If VREAD  VDD , then the output registers a voltage VBIT = VDD denoting the binary state “1,” but if VREAD ≥ VDD , then the output terminal shorts to ground, generating VBIT = 0 denoting the binary state “0.” But how does this translate into a permanent memory on a chip? Let us examine the 4-bit ROM diagrammed in Fig. TF10-2(b). In this case, some bits simply do not have transistors; VBIT2 , for example, is permanently connected to VDD via a resistor. This may seem trivial,

Figure TF10-1: Integrated circuit die photo of a Micron MT4C1024 220 -bit DRAM chip. Die size is 8.662 mm × 3.969 mm. (Courtesy of ZeptoBars.)

“book” — 2015/5/4 — 7:10 — page 204 — #22

204

TECHNOLOGY BRIEF 10: COMPUTER MEMORY CIRCUITS

VDD (dc voltage source) R VBIT VREAD

(a) 1-bit ROM VDD (dc voltage source)

R VBIT1

R VBIT2

R VBIT3

R VBIT4

VREAD

(b) 4-bit ROM Figure TF10-2: (a) 1-bit ROM that uses a MOSFET transistor, and (b) 4-bit ROM configured to store the sequence [0100], whose decimal value is 4.

Random-Access Memories (RAMs)

categories: static RAMs and dynamic RAMs (DRAMs). Because RAMs lose the state of their bits if the power is removed, they are termed volatile memories. Static RAMs not only can be read from and written to, but also do not forget their state as long as power is supplied. These circuits also are composed of transistors, but each single bit in a modern static RAM consists of four transistors wired up in a bi-stable circuit (the explanation of which we will leave to your intermediate digital components classes!). Dynamic RAMs, on the other hand, are illustrated more easily. Dynamic RAMs usually hold more bits per area than static RAMs, but they need to be refreshed constantly (even when power is supplied continuously to the chip).

RAMs are a class of memories that can be read to and written from constantly. RAMs generally fall into two

Figure TF10-3 shows a simple one-transistor dynamic RAM. Again, we will treat the transistor as we did in

but this specific 4-bit memory configuration always stores the value [0100]. In this same way, thousands of such components can be strung together in rows and columns in N × N arrays. As long as a power supply of voltage VDD is connected to the circuit, the memory will report its contents to an external circuit as [0100]. Importantly, even if you remove power altogether, the values are not lost; as soon as you add power back to the chip, the same values appear again (i.e., you would have to break the chip to make it forget what it is storing!). Because of the permanency of this data, these memories also often are called nonvolatile memories (NVM).

“book” — 2015/5/4 — 7:10 — page 205 — #23

TECHNOLOGY BRIEF 10: COMPUTER MEMORY CIRCUITS

205

Vcolumn To other cells

Vrow N1

C

To other cells Figure TF10-3: 1-bit DRAM cell.

Section 4-11. Note that if we make VROW > VDD , then the transistor will conduct and the capacitor C will start charging to whatever value we select for VCOLUMN . When writing a bit, VCOLUMN usually is set at either 0 (GND) or 1 (VDD ). We can calculate how long this chargingup process will require, because we know the value of C and the transistor’s current gain g (see Section 5-7). When the capacitor is charged to VDD , a value of 1 is stored in the DRAM. Had we applied instead a value of zero volts to VCOLUMN , the transistor would have discharged to ground (instead of charged to VDD ) and the bit would have a value of 0. However, note that unlike the ROM, the state of the bit is not “hardwired.” That is, if even tiny leakage currents were to flow through the transistor when it is not on (that is, when VROW < VDD ), then charge will constantly leak away and the voltage of the transistor will drop slowly with time. After a short time (on the order of a few milliseconds in the dynamic RAM in your computer), the capacitor will have irrecoverably lost its value. How is that mitigated? Well, it turns out that a modern memory will read and then re-write every one of its (several billion) bits every 64 milliseconds to keep them refreshed! Because each bit is so simple (one transistor and one capacitor), it is possible to manufacture DRAMs with very high memory densities (which is why 1-Gbit DRAMs are now available in packages of reasonable size). Other variations of DRAMs also exist whose architectures deviate slightly from the previous model—at either the transistor or system level. Synchronous Graphics RAM (SGRAM), for example, is a DRAM modified for use with graphics adaptors; Double Data Rate 4 RAM (DDR4RAM) is a fourth-generation enhancement over DRAM which allows for faster clock speeds and lower operating voltages.

Advanced Memories Several substantially different technologies are emerging that likely will change the market landscape—just as Flash memories revolutionized portable memory (like your USB memory stick). Apart from the drive to increase storage density and access speed, one of the principal drivers in today’s memory research is the development of non-volatile memories that do not degrade over time (unlike Flash). The Ferroelectric RAM (FeRAM) is the first of these technologies to enter mainstream production; FeRAM replaces the capacitor in DRAM (Fig. TF10-3) with a ferroelectric capacitor that can hold the binary state even with power removed. While FeRAM can be faster than Flash memories, FeRAM densities are still much smaller than modern Flash (and Flash densities continue to increase rapidly). FeRAM currently is used in niche applications where the increased speed is important. Magnetoresistive RAM (MRAM) is another emerging technology, currently commercialized by Everspin Technologies (spun out from Freescale Semiconductor), which relies on magnetic plates to store bits of data. In MRAM, each cell is composed of two ferromagnetic plates separated by an insulator. The storage and retrieval of bits occurs by manipulation of the magnetic polarization of the plates with associated circuits. Like FeRAM, MRAM currently is overshadowed by Flash memories, but improvements in density, speed, and fabrication methods may make it a viable alternative in the mainstream consumer market in the future. Even more speculative is the idea of using single carbon nanotubes to store binary bits by changing their configuration electronically; this technology is currently known as Nano RAM (NRAM).

“book” — 2015/5/4 — 7:10 — page 206 — #24

206

CHAPTER 4

υp ip = 0

Rs1 Rs2 υ1

+ _

υ2

υn

+ _

tion/summation is done must keep each individual stage from exceeding +/ − Vcc .

+ υo

in = 0 −

R1 R2

1

s2

2

G1 =

υ2

R1 + R2 G1 = R2

2

(

s1

)(

s2

Rs1 Rs1 + Rs2

• Sensitivity when adding large and small values. Care is typically taken to add values that are similar in magnitude, so amplification is typically done prior to summation if two values have significantly different magnitudes. • Other considerations . . . Concept Question 4-12: What type of op-amp circuits (inverting, noninverting, and others) might one use to perform the operation υo = G1υ1 +G2υ2 with G1 and G2 both positive? (See )

( R R+ R ) ( R R+ R )

υ1

OPERATIONAL AMPLIFIERS

+

)

υo = G1υ1 + G2υ2

Concept Question 4-13: What is an inverting adder?

(See

)

Exercise 4-5: The circuit shown in Fig. 4-14(a) is to be used to perform the operation

υo = 3υ1 + 6υ2 .

Figure 4-15: Noninverting summer.

If R1 = 1.2 k, Rs2 = 2 k, and Rf2 = 4 k, select values for R2 and Rf1 so as to realize the desired result. (b) υo = (−4υ1 −7υ2 )(−1): Multiply υ1 by −4 and υ2 by −7 and add them together (inverting summing amplifier with gains of −4 and −7), and then multiply the result by −1 (inverting amplifier with a gain of −1). (c) υo = (4υ1 + 7υ2 ): Multiply υ1 by 4 and υ2 by 7 and add them together (noninverting summing amplifier with gains of 4 and 7). (d) υo = [(2υ1 )+(3.5υ2 )]×2: Multiply υ1 by 2 (noninverting amplifier with a gain of 2) and υ2 by 3.5 (noninverting amplifier with a gain of 3.5), and then add them (noninverting summer with a gain of 2). Why might you choose one of these systems over another? There are several reasons: • To minimize the number of op amps (option c) • To meet gain limitations. An inverting amplifier can have a gain of less than 1, but a noninverting amplifier cannot. • To avoid saturation. The output voltage of any individual stage is limited by its Vcc . The order in which multiplica-

Answer: Rf1 = 1.8 k, R2 = 600 . (See

4-6

)

Difference Amplifier

When an input signal υ2 is connected to terminal υp of a noninverting amplifier circuit, the output is a scaled version of υ2 . A similar outcome is generated by an inverting amplifier circuit when an input voltage υ1 is connected to the op amp’s υn terminal, except that in addition to scaling υ1 its polarity is reversed as well. The difference amplifier circuit combines these two functions to perform subtraction. In the difference-amplifier circuit of Fig. 4-16(a), the input signals are υ1 and υ2 , R2 is the feedback resistance, R1 is the source resistance of υ1 , and resistances R3 and R4 serve to control the scaling factor (gain) of υ2 . To obtain an expression that relates the output voltage υo to the inputs υ1 and υ2 , we apply KCL at nodes υn and υp . At υn , i1 + i2 + in = 0, which is equivalent to υn − υo υn − υ 1 + + in = 0 R1 R2

(node υn ).

(4.38)

“book” — 2015/5/4 — 7:10 — page 207 — #25

4-6

DIFFERENCE AMPLIFIER

207 where the scale factors (gains) are given by

Difference Amplifier R2 R1

υ1

R3

+ _

i1

i2 in = 0

i3

υn i = 0 p υp

υ2

+ _

 G2 =

_

υo

+

i4

G2 =

υ1

(

R4 R3 + R4

G1 = −

)(

R1 + R2 R1

R1 + R2 R1

 (4.42a)

 G1 = −

R2 R1

 .

(4.42b)

According to Fig. 4-16(b) which is a block-diagram representation of the difference amplifier circuit:

(a) Difference circuit

υ2



and

RL

R4

R4 R3 + R 4

 The difference amplifier scales υ2 by positive gain G2 , υ1 by negative gain G1 and adds them together. 

) +

R2 R1

For the difference amplifier to function as a subtraction circuit with equal gain, its resistors have to be interrelated by

υo = G1υ1 + G2υ2

R2 R3 = R1 R4 ,

(4.43)

(b) Block diagram in which case Eq. (4.41) reduces to Figure 4-16: Difference-amplifier circuit.  At υp , i3 + i4 + ip = 0, or

υo =

υp υp − υ2 + + ip = 0 R3 R4

(node υp ).

υo =

R4 R3 + R 4



R1 + R2 R1



 υ2 −

R2 R1

(υ2 − υ1 )

(equal gain).

(4.44)

Exact subtraction with no scaling requires that R1 = R2 . Exercise 4-6: The difference-amplifier circuit of Fig. 4-16

 υ1 ,

is used to realize the operation υo = (6υ2 − 2) V.

(4.40) which can be cast in the form υo = G2 υ2 + G1 υ1 ,



(4.39)

Upon imposing the ideal op-amp constraints ip = in = 0 and υp = υn , we end up with 

R2 R1

(4.41)

Given that R3 = 5 k, R4 = 6 k, and R2 = 20 k, specify values for υ1 and R1. Answer: υ1 = 0.2 V, R1 = 2 k. (See

C3

)

“book” — 2015/5/4 — 7:10 — page 208 — #26

208

CHAPTER 4

4-7 Voltage Follower/Buffer 4-7.1

Rs

No Buffer

In electronic circuits, we often need to incorporate the functionality of a relatively simple (but important) circuit that serves to isolate the input source from variations in the load resistance RL . Such a circuit is called a voltage follower, buffer, or unity gain amplifier. To appreciate the utility of the voltage follower, let us first examine the circuit shown in Fig. 4-17(a). A source input circuit represented by its Th´evenin equivalent (υs , Rs ), is connected to a load RL . The output voltage is

(without voltage follower),

(4.45)

which obviously is dependent on both Rs and RL , so if the load resistance RL changes, so will the output voltage υo .

υo

+ _

υs

υ s RL υo = Rs + R L

OPERATIONAL AMPLIFIERS

No buffer

RL

Source circuit

Load

(a) Source circuit connected directly to a load

υp ip = 0

Rs + _

υs

υn in = 0

Buffer

+

υo

_

RL

Source circuit

Load

(b) Source circuit separated by a buffer Figure 4-17: The voltage follower provides no voltage gain (υo = υs ), but it insulates the input circuit from the load.

4-7.2 With Op-Amp Buffer In contrast, when the op-amp voltage follower circuit shown in Fig. 4-17(b) is inserted in between the source circuit and the load, the output voltage becomes completely independent of both Rs and RL . Because ip = 0, it follows that υp = υs . Furthermore, in view of the op-amp constraint υp = υn and because the output node is connected directly to υn , it follows that υo = υp = υs

(with voltage follower),

 When designing and building a multistage circuit, designers usually insert buffers between adjacent stages, which allows them to design each stage separately and then cascade them all together with buffers in between them. 

(4.46)

and this is true regardless of the values of Rs and RL (excluding Rs = open circuit and/or RL = short circuit, either of which would invalidate the entire circuit). Thus:

4-7.3

Input-Output Resistance

When is a buffer needed? Consider again the circuit in Fig. 4-17(a). Let us examine υo for various values of Rs and RL . Rs (k)

 The output of the voltage follower follows the input signal while remaining immune to changes in RL because it has a high input resistance and low output resistance. 

A circuit that offers this type of protection is often called a buffer.

1 1 1 1

RL () 100 1 10 100

υo (V) 0.09 0.5 0.91 0.99

% change 91% 50% 9% 1%

Buffer needed? Yes Yes Probably No

If Rs < RL , or even if Rs ≈ RL , there is a substantial difference between υo and υs . This is overloading the circuit, which we

“book” — 2015/5/4 — 7:10 — page 209 — #27

4-8

OP-AMP SIGNAL-PROCESSING CIRCUITS

209

typically just call loading. Substantial current is drawn from the source, and the voltage is decreased as a result. To prevent this, a buffer is needed. But if Rs  RL , the change is minimal, and the circuit does not require a buffer. An additional interesting aspect of buffering has to do with where the current is coming from and where it is going to in the circuit. In Fig. 4-17(a), the current is coming from the source and going to the load. Excess current is being drawn, and the circuit is (over)loaded, thus reducing the output voltage υo . In Fig. 4-17(b), the current is not coming from the source, but it is going to the load. Where is it coming from? The answer is that it is coming from the output of the buffer, extracted from the power supply voltage Vcc that powers the op amp in the buffer.

Concept Question 4-14: What is the function of a voltage

follower, and why is it called a “buffer”? (See

)

These circuits can be used in various combinations to realize specific signal-processing operations. We note that the input-output transfer functions are independent of the load resistance RL that may be connected between the output terminal υo and ground. In the case of the noninverting amplifier, the transfer function is also independent of the source resistance Rs .  When cascading multiple stages of op-amp circuits in series, care must be exercised to ensure that none of the op amps is driven into saturation by the cumulative gain of the multiple stages.  When analyzing circuits that involve op amps, whether in configurations similar to or different from those we encountered so far in this chapter, the basic rules to remember are as follows:

Basic Rules of Op-Amp Circuits Concept Question 4-15: How much voltage gain is

provided by the voltage follower? (See

)

Exercise 4-7: Express υo in terms of υ1 , υ2 , and υ3 for

the circuit in Fig. E4.7.

3 kΩ υ1

0.5 kΩ

υ2

1 kΩ

υ3

2 kΩ

_ +

10 kΩ 5 kΩ

_ +

υo

(2) The op amp will operate in the linear range so long as |υo | < |Vcc |. (3) The ideal op-amp model assumes that the source resistance Rs (connected to terminals υp or υn ) is much smaller than the op-amp input resistance Ri (which usually is no less than 10 M), and the load resistance RL is much larger than the op-amp output resistance Ro (which is on the order of tens of ohms). (4) The ideal op-amp constraints are ip = in = 0 and υp = υn .

Figure E4.7 Answer: υo = 12υ1 + 6υ2 + 3υ3. (See

(1) KCL and KVL always apply everywhere in the circuit, but KCL is inapplicable at the output node when applying the ideal op-amp model. All other circuit-analysis tools can be applied to op-amp circuits.

C3

) Example 4-4: Block-Diagram Representation

4-8 Op-Amp Signal-Processing Circuits Table 4-3 provides a summary of the op-amp circuits we have considered thus far, together with their functional characteristics in the form of block-diagram representations.

Generate a block-diagram representation for the circuit shown in Fig. 4-18(a). Solution: The first op amp is an inverting amplifier (Table 4-3(b)) with a dc input voltage υ1 = 0.42 V. Its circuit gain Gi (with the subscript added to denote “inverting amp”) is Gi = −

30K = −3, 10K

“book” — 2015/5/4 — 7:10 — page 210 — #28

210

CHAPTER 4

OPERATIONAL AMPLIFIERS

Table 4-3: Summary of op-amp circuits. Op-Amp Circuit (a)

υs

+

Rs

G=

υs

R1 + R2 R2

υo = Gυs

R1

R2

υs

υo

υ



(b)

Block Diagram

Noninverting Amp (υo independent of Rs)

Rs

Rf

υs



G=−

υo

+

Rf Rs

υo = Gυs

Inverting Amp (c) υ1

R1 R2

υ2

R3

υ3

Rf

υ1

G1 = − Rf /R1



υ2

G2 = − Rf /R2

υ3

G3 = − Rf /R3

υo

+

+

υo = G1υ1 + G2υ2 + G3υ3

Inverting Summing Amp (d) υ1

R2

R1

G1 = −



R3

υ2

υ1 υo

+

R4

υ2

G2 =

(

R2 R1 R1 + R2 R1

)(

R4 R3 + R4

+

υo = G1υ1 + G2υ2

)

Subtracting Amp (e)

υs

Rs

+

G=1

υs

υo



υo = υs

RL Voltage Follower / Buffer (υo independent of Rs and RL) υp ip = 0

Rs1

(f)

Rs2 υ1

+ _

υ2

+ _

υn

+ −

R1

1

υ2

R1 + R2 G2 = R2

R2 Noninverting Summing Amp

s2

2

G1 =

υo

in = 0

( R R+ R ) ( R R+ R )

υ1

2

(

s1

)(

s2

Rs1 Rs1 + Rs2

+

)

υo = G1υ1 + G2υ2

“book” — 2015/5/4 — 7:10 — page 211 — #29

4-8

OP-AMP SIGNAL-PROCESSING CIRCUITS

211 is open to the outside air, is P . When at sea level, P = P0 , so the membrane assumes a flat shape and the two capacitances are equal. Since atmospheric pressure decreases with elevation, a rise in altitude results in a change in the pressure P in the upper chamber, causing the membrane to bend upwards (Fig. 4-19(b)), thereby changing the capacitances of the two capacitors. The sensor measures a voltage υs that is proportional to the change in capacitance. Based on measurements of υs as a function of h, the data was found to exhibit an approximately linear variation given by

and its output is υo1 = Gi υ1 = −3(0.42) = −1.26 V. The second op amp is a difference amplifier. Using Table 4-3(d), the gains of its positive and negative channels are  R1 + R2 R1    2K 10K + 20K = =2 1K + 2K 10K 

G2 =

R4 R3 + R 4



υs = 2 + 0.2h

and G1 = −

R2 20K = −2. =− R1 10K

υo = G2 υ2 + G1 υo1 = 2υ2 − 2(−1.26) = (2υ2 + 2.52) V.

Solution: Based on the given information, the sensor voltage υs will serve as the input to the circuit we are asked to design, and the output υo will represent the height elevation h. We therefore need a circuit that can perform the operation

Example 4-5: Elevation Sensor

υo = h =

A hand-held elevation sensor uses a pair of capacitors separated by a flexible metallic membrane (Fig. 4-19(a)) to measure the height h above sea level. The lower chamber in Fig. 4-19(a) is sealed, and its pressure is P0 , which is the standard atmospheric pressure at sea level. The pressure in the upper chamber, which

υ1

10 kΩ

1 2 υs − = 5υs − 10, 0.2 0.2

where we have inverted Eq. (4.47) to solve for h in terms of υs . The functional form of Eq. (4.48) indicates that we have

υo1

++

10 kΩ − Op Amp 2

1 kΩ υ2

+ _

+

2 kΩ

(a) Circuit −1.26 V

(b) Block diagram

−3

(4.48)

20 kΩ

30 kΩ −Op Amp 1

0.42 V

(4.47)

where h is in km. The sensor is designed to operate over the range 0 ≤ h ≤ 10 km. Design a circuit whose output voltage υo (in volts) is an exact indicator of the height h (in km).

Hence,

0.42 V

(V),

υo1 υ2

−2 2

2.52 V

2υ2

+

υo = (2υ2 + 2.52) V

Figure 4-18: Block-diagram representation (Example 4-4).

υo

+

“book” — 2015/5/4 — 7:10 — page 212 — #30

212

CHAPTER 4

OPERATIONAL AMPLIFIERS

Equation (4.49) can be made to implement Eq. (4.48) if we select the following

Air

(a) υs = υ2

Metal plate

1

Flexible metal membrane

2

(c) values for R1 through R4 that simultaneously satisfy the conditions    R2 R1 + R2 R4 = 5. = 10 and R1 R3 + R 4 R1

C2

P0 Metal plate (a)

(b) υ1 as a dc voltage source such that (R2 /R1 )υ1 = 10 V, which can be satisfied by arbitrarily selecting υ1 = 1 V and (R2 /R1 ) = 10

C1

P

3 Pressure sensor

A possible set of values that meets these conditions is

1 1

P P < P0

C1

2

2

P0

C2

Sensor

3

Capacitances R2 R1

+ −

υ1 = 1 V

υn R3

υp



R2 = 20 k,

R3 = 10 k,

R4 = 8.33 k.

Before we conclude the design, we should check to make sure that the op amp will operate in its linear range over the full range of operation of the sensor. According to Eq. (4.47), as h varies from zero to 10 km, υs varies from 2 V to 4 V. The corresponding range of variation of υo , from Eq. (4.48), is from zero to 10 V. Hence, we should choose an op amp designed to function with a dc supply voltage Vcc that exceeds 10 V.

3 (b)

R1 = 2 k,

υo

+

Example 4-6: Circuit with Multiple Op Amps

Sensor

(c)

υ2 = υs

R4

R1 = 2 kΩ R2 = 20 kΩ R3 = 10 kΩ R4 = 8.33 kΩ

Circuit realization

Figure 4-19: Design of a circuit for the pressure sensor of Example 4-5 with P0 = pressure at sea level and P = pressure at height h.

only one active (variable) input, namely υs , which we need to amplify by a factor of 5, but we also need to subtract 10 V from it. There are multiple circuit configurations that can achieve the desired operation, including the subtractor circuit shown in Table 4-3(d) and in Fig. 4-19(c). According to Eq. (4.40), the output of the difference amplifier is given by      R1 + R2 R4 R2 υo = υ2 − υ1 . (4.49) R3 + R 4 R1 R1

Relate the output voltage υo to the input voltages υ1 and υ2 of the circuit in Fig. 4-20. Solution: By comparing the circuit connections surrounding the four op amps with those given in Table 4-3, we recognize op amps 1 and 2 as noninverting amplifiers (sources υ1 and υ2 are connected to + input terminals), op amp 3 as an inverting amplifier with a gain of −1 (equal input and feedback resistors R4 ), and op amp 4 as an inverting summing amplifier (Table 4-3(b)) with equal gain (same input resistances R6 at summing point). We start by examining the pair of input op amps. Because they are not among the standard configurations in Table 4-3, we will use KVL/KCL to evaluate them. For op amp 1, υp1 = υ1 and υp1 = υn1 (op-amp voltage constraint). Hence, υa = υn1 = υ1 . Similarly, for op amp 2, υb = υn2 = υ2 .

“book” — 2015/5/4 — 7:10 — page 213 — #31

4-8

υ1

OP-AMP SIGNAL-PROCESSING CIRCUITS

+ _

υp1 υn1

+

213

R6

υo1

Op Amp 1 −

R1

in1 = 0

R4

υa in2 = 0 υn2 υp2 υ2

+ _

i2

R3

R4

R5 υo

+

− Op Amp 3

+

υo2

+

υ'o2

R2 υb

− Op Amp 2

R6

− Op Amp 4

Inverting Summing Amp

Inverting Amp Noninverting Amps Figure 4-20: Example 4-6.

Since in1 = in2 = 0 (op-amp current constraint), i2 =

Example 4-7: Interesting Op-Amp Circuit

υb − υa υ2 − υ1 = , R2 R2

Generate a plot for iL at the output side of the circuit shown in Fig. 4-21(a) versus υs , covering the full linear range of υs .

and υo2 − υo1 = i2 (R1 + R2 + R3 )   R1 + R2 + R3 = (υ2 − υ1 ). R2

Solution: This circuit is not one of the standard op-amp configurations in Table 4-3, so we need to analyze it using KVL/KCL. At node υn , KCL gives (4.50)

Op amp 3 is a standard inverting amplifier, so we can use Table 4-3(c) to obtain υo 2 = −



R4 R4

υ n − υo υn + = 0, 2k 6k which leads to

 υo2 = −υo2 .

υo = 4υn . At node υp , KCL gives

Op amp 4 is an inverting summing amplifier (Table 4-3(c)) with output R5 (υo + υo 2 ) R6 1 R5 = − (υo1 − υo2 ) R6   R5 R1 + R 2 + R 3 = (υo − υo1 ) = R5 (υ2 − υ1 ). R6 2 R6 R2 (4.51)

υo = −

υp − (υs − 0.5) = 0, 2k which leads to υp = υs + 0.5. By imposing the op-amp constraint υp = υn , we have υo = 4υn = 4(υs + 0.5) = 4υs + 2.

“book” — 2015/5/4 — 7:10 — page 214 — #32

214

CHAPTER 4

OPERATIONAL AMPLIFIERS

At the output side, iL =

4υs + 2 − 4 υo − 4 = = (4υs − 2) mA. 1k 1k

2 kΩ

10 = 4υs + 2,

+ _

or υs = 2 V,

υn υp

(υs + 0.5)

For υo = Vcc = 10 V,

6 kΩ

in = 0

2 kΩ

ip = 0

_ +

0.5 V

υs

υo iL

Vcc = 10 V RL 1 kΩ + _ 4V

and for υo = −Vcc = −10 V,

(a) Circuit −10 = 4υs + 2,

or υs = −3 V.

iL(mA)

Hence, linear range of υs is −3 V ≤ υs ≤ 2 V

15

(linear range).

12

Figure 4-21(b) displays a plot of iL versus υs over the latter’s linear range. Note that the linear range is not symmetrical.

9 6 3

4-9 Instrumentation Amplifier −4 An electric sensor is a circuit used to measure a physical quantity, such as distance, motion, temperature, pressure, or humidity. In some applications, the intent is not to measure the magnitude of a certain quantity, but rather to sense small deviations from a nominal value. 

−3

−2

−1

−3

2

3

4

υs (V)

−6 −9 −12 −15

For example, if the temperature in a room is to be maintained at 20 ◦ C, the functional goal of the temperature sensor is to measure the difference between the room temperature T and the reference temperature T0 = 20 ◦ C and then to activate an air conditioning or heating unit if the deviation exceeds a certain prespecified threshold. Let us assume the threshold is 0.1 ◦ C. Instead of requiring the sensor to be able to measure T with an absolute accuracy of no less than 0.1 ◦ C, an alternative approach would be to design the sensor to measure υ = υ2 − υ1 , where υ2 is the voltage output of a thermocouple circuit responding to the room temperature T and υ1 is the voltage corresponding to what a calibrated thermocouple would measure when T0 = 20 ◦ C. Thus, the sensor is designed to measure the deviation of T from T0 , rather than T itself, with an absolute accuracy of no less than 0.1 ◦ C. The advantage of such an approach is that the signal is now υ, which is more than two orders of magnitude smaller than υ2 . A circuit with a precision of 10 percent is not good enough for measuring υ2 , but it is plenty good for measuring υ.

1

−14

(b) iL − υs transfer plot Figure 4-21: Circuit for Example 4-7.

To appreciate the advantage of the differential measurement approach over the direct measurement approach, consider the two system configurations represented in Fig. 4-22. (a) Direct Measurement Approach In the configuration depicted in Fig. 4-22(a), input voltage υ2 represents the voltage across a thermistor used to measure the temperature T in a house. The voltage is related to T by υ2 = 0.01T , with T in ◦ C. The application circuit has a gain of 100 and a measurement precision of ±1% of the amplified output. Thus, υo = (100 ± 1)υ2 = (100 ± 1) × 0.01T = T ± 0.01T .

“book” — 2015/5/4 — 7:10 — page 215 — #33

4-9

INSTRUMENTATION AMPLIFIER

Thermistor υ2 = 0.01T

υ2

G = 100 ± 1% of υo

υo = (G ± 1)υ2 = T ± 0.01T. For T = 21 ˚C, υo = (21 ± 0.21) ˚C.

(a) Direct measurement

Thermistor υ2 = 0.01T

υ2

G = 100

+ _ ± 1% of υo

υ1 = 0.2 V

Fixed reference temperature = 20 ˚C

υo = G(υ2 − υ1) ± 1% of G(υ2 − υ1) = (T − 20) ± 0.01(T − 20). For T = 21 ˚C, υo = (1 ± 0.01) ˚C.

Much better measurement uncertainty

(b) Differential measurement Figure 4-22: Comparison of direct and differential measurement uncertainties.

215 provided by the direct measurement system, but with an associated precision on the order of 20 times better (±0.01 ◦ C compared with ±0.21 ◦ C for the direct measurement system).  The instrumentation amplifier is perfectly suited for detecting and amplifying a small signal deviation when superimposed on one or the other of two much larger (and otherwise identical) signals.  An instrumentation amplifier consists of three op amps, as shown in Fig. 4-23. The circuit configuration for the first two is the same as the one we examined earlier in connection with Example 4-6. According to Eq. (4.50), the voltage difference between the outputs of op amps 1 and 2 is   R1 + R2 + R3 υo2 − υo1 = (υ2 − υ1 ) = G1 (υ2 − υ1 ), R2 (4.52) where G1 is the circuit gain of the first stage (which includes op amps 1 and 2) and is given by G1 =

R1 + R2 + R3 . R2

(4.53)

The third op amp is a difference amplifier that amplifies (υo2 − υo1 ) by a gain factor G2 given by If T = 21 ◦ C, the output registers 21 ◦ C, and the associated precision is 0.21 ◦ C. (b) Differential Measurement Approach The differential system in Fig. 4-22(b) also uses υ2 to measure T , but it also uses a fixed voltage υ1 at the negative terminal, with υ1 set at the desired reference temperature of 20 ◦ C. Hence, υ1 = 0.2 V. The differential output is given by υo = 100(υ2 − υ1 ) ± (υ2 − υ1 ) = 100(υ2 − 0.2) ± (υ2 − 0.2) = 100(0.01T − 0.2) ± (0.01T − 0.2) = (T − 20) ± 0.01(T − 20). If T = 21 ◦ C, υo = (1 ± 0.01) ◦ C. In the differential system, υo measures the deviation from the reference temperature of 20 ◦ C, which is the same information

G2 =

R4 . R5

(4.54)

Hence,  υo = G2 G1 (υ2 − υ1 ) =

R4 R5



R1 + R2 + R3 R2

 (υ2 − υ1 ).

(4.55) To simplify the circuit, and improve precision, all resistors— with the exception of R2 —often are chosen to be identical in design and construction, thereby minimizing deviations between their resistances. If we set R1 = R3 = R4 = R5 = R in Eq. (4.55), the expression for υo reduces to   2R (υ2 − υ1 ). υo = 1 + R2

(4.56)

In that case, R2 becomes the gain-control resistance of the circuit; its value (relative to R) sets the gain. If the expected signal deviation (υ2 − υ1 ) is on the order of microvolts to millivolts, the instrumentation amplifier is designed to have an overall gain that would amplify the signal to the order of volts.

“book” — 2015/5/4 — 7:10 — page 216 — #34

216

CHAPTER 4

OPERATIONAL AMPLIFIERS

Instrumentation Amplifier +

υ1

Gain control

Op Amp 1 − R1

υo1

R5

R3

υo2

R5

R2

υ2

R4 − Op Amp 3

υo

+

− Op Amp 2

R4

υ1 υ2

+

G=

( RR ) ( R + RR + R ) 4

1

5

2

3

2

G(υ2 − υ1) Figure 4-23: Instrumentation-amplifier circuit.

 The instrumentation amplifier is a high-sensitivity, high-gain, deviation sensor. Several semiconductor manufacturers offer instrumentation-amplifier circuits in the form of integrated packages.  Concept Question 4-16: When designing a multistage

op-amp circuit, what should the design engineer do to insure that none of the op amps is driven into saturation? (See ) Concept Question 4-17: If the goal is to measure small

deviations between a pair of input signals, what is the advantage of using an instrumentation amplifier over using a difference amplifier? (See )

Exercise 4-8: To monitor brain activity, an instrumentation-amplifier sensor uses a pair of needlelike probes inserted at different locations in the brain to measure the voltage difference between them. If the circuit is of the type shown in Fig. 4-23 with R1 = R3 = R4 = R5 = R = 50 k, Vcc = 12 V, and the maximum magnitude of the voltage difference that the brain is likely to exhibit is 3 mV, what should R2 be to maximize the sensitivity of the brain sensor? Answer: R2 = 25 . (See

C

)

4-10

Digital-to-Analog Converters (DAC)

 A digital-to-analog converter (DAC) is a circuit that transforms a digital sequence presented to its input into an analog output voltage whose magnitude is proportional to the decimal value of the input signal.  An n-bit digital signal is described by the sequence [V1 V2 V3 . . . Vn ], where V1 is called the most significant bit (MSB) and Vn is the least significant bit (LSB). Voltages V1 through Vn can each assume only two possible states—either a 0 or a 1. When a bit is in the 1 state, its decimal value is 2m , where m depends on the location of that bit in the sequence. For the most significant bit (V1 ), its decimal value is 2(n−1) ; for V2 it is 2(n−2) ; and so on. The decimal value of the least significant bit is 2n−n = 20 = 1, when that bit is in state 1. Any bit in state 0 has a decimal value of 0. Table 4-4 illustrates the correspondence between the binary sequences of a 4-bit digital signal and their decimal values. The binary sequences start at [0000] and end at [1111], representing 16 decimal values extending from 0 to 15 and inclusive of both ends. To do so, the DAC in Fig. 4-24 has to sum V1 to Vn after weighting each by a factor equal to its decimal value. Thus, for a 4-bit digital sequence, for example, the output voltage of the DAC has to be related to the input by Vout = G(24−1 V1 + 24−2 V2 + 24−3 V3 + 24−4 V4 ) = G(8V1 + 4V2 + 2V3 + V4 ),

(4.57)

“book” — 2015/5/4 — 7:10 — page 217 — #35

4-10

DIGITAL-TO-ANALOG CONVERTERS (DAC)

217

V1

MSB

V2

n-bit digital input signal [V1V2KVn]

M

DAC

Vout = G(2n −1V1 + 2n −2V2 + L + 2Vn −1 + Vn)

Vn

LSB Figure 4-24: A digital-to-analog converter transforms a digital signal into an analog voltage proportional to the decimal value of the digital sequence.

to either Table 4-3(c) or Eq. (4.34) yields

Table 4-4: Correspondence between binary sequence and decimal value for a 4-bit digital signal and output of a DAC with G = −0.5.

V1 V2 V3 V4

Decimal Value

0000 0001 0010 0011 0100 0101 0110 0111 1000 1001 1010 1011 1100 1101 1110 1111

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

DAC Output (V) 0 −0.5 −1 −1.5 −2 −2.5 −3 −3.5 −4 −4.5 −5 −5.5 −6 −6.5 −7 −7.5

where G is a scale factor that has no influence on the relative weights of the four terms. The magnitude of G is selected to suit the range of the output voltage. If the input is a 3-bit sequence whose range of decimal values extends from 0 to 7, one might design the circuit so that G = 1, because in that case, the maximum output voltage is 7 V, which is below Vcc for most op amps. For digital signals with longer sequences, G needs to be smaller than 1 in order to avoid saturating the op amp. The weighted-sum operation of a DAC can be realized by many different signal-processing circuits. A rather straightforward implementation is shown in Fig. 4-25, where an inverting summer (Table 4-3(c)) uses the ratios of Rf to the individual resistances to realize the necessary weights, and the positions of the switches determine the 0/1 states of the 4 bits. Reference

Rf Rf Rf Rf V1 − V2 − V3 − V4 R 2R 4R 8R −Rf = (4.58) (8V1 + 4V2 + 2V3 + V4 ), 8R

Vout = −

which satisfies the relative weights given in Eq. (4.57). Also, in this case, G=−

Rf . 8R

(4.59)

For [V1 V2 V3 V4 ] = [1111], Vout = 15G. By selecting G = −0.5 (corresponding to Rf = 4R), the output will vary from 0 to −7.5.

Example 4-8: R–2R Ladder

The circuit in Fig. 4-26(a) offers an alternative approach to realizing digital-to-analog conversion of a 4-bit signal. It is called an R–2R ladder, because all of the resistors of its input circuit have values of R or 2R, thereby limiting the input resistance seen by the dc source to a 2 : 1 range no matter how many bits are contained in the digital sequence. This is in contrast with the DAC of Fig. 4-25, whose inputresistance range is dependent on the number of bits; 8 : 1 for a 4-bit converter, and 128 : 1 for an 8-bit converter. Additionally, circuit performance and precision depend on resistor tolerance and are superior when fewer groups of resistors are involved in the input circuit. Resistors fabricated in the same production process are likely to exhibit less variability among them than resistors fabricated by different processes. Show that the R–2R ladder in Fig. 4-26(a) does indeed provide the appropriate weighting for a 4-bit DAC. If R = 2 k and Vcc = 10 V, what is the maximum realistic value that Rf can have?

“book” — 2015/5/4 — 7:10 — page 218 — #36

218

CHAPTER 4

OPERATIONAL AMPLIFIERS

Rf

LSB

8R

4R

2R

V4

V3

V2

0

1

0

1

0

R V1

MSB 1

0

1

+ −



+

+

Vout

Inverting Summing Amp

1V



Figure 4-25: Circuit implementation of a DAC.

R 2R

R

Rf

R

2R

2R

2R

2R

V4 LSB

V3

V2

V1 MSB



+

1

+

Vout

0

1

0

1

0

1

0

+ −

Inverting Summing Amp

1V

− R−2R ladder network

(a)

Rf

RTh (b)

' Thevenin equivalent circuit

V2 V3 V1 V4 VTh = + + + 2 4 8 16 RTh = R

+ −

VTh



+

+

Vout

Figure 4-26: R–2R ladder digital-to-analog converter.



“book” — 2015/5/4 — 7:10 — page 219 — #37

4-11 THE MOSFET AS A VOLTAGE-CONTROLLED CURRENT SOURCE Solution: Even though we know that (depending on the positions of the switches) V1 to V4 can each assume only 2 binary values, namely 0 or 1 V, let us treat V1 to V4 as dc power supplies and apply multiple iterations of voltage-current transformations (starting on the left with the LSB) to arrive at the Th´evenin equivalent circuit at the input side of the op amp. The result of such a transformation process is shown in Fig. 4-26(b), in which VTh =

V1 V2 V3 V4 + + + 2 4 8 16

219

Drain (D)

D

Insulator

+ G

IDS

+ VGS _

Gate (G)

Source (S) (a) MOSFET symbol

VDS _

S (b) Voltages

(4.60a) Figure 4-27: MOSFET symbol and voltage designations.

and RTh = R.

(4.60b)

Consequently, Exercise 4-9: A 3-bit DAC uses an R–2R ladder design with R = 3 k and Rf = 24 k. If Vcc = 10 V, write an expression for Vout and evaluate it for [V1 V2 V3 ] = [111].

Rf VTh RTh   Rf V1 V2 V3 V4 =− + + + R 2 4 8 16

Vout = −

Rf =− (8V1 + 4V2 + 2V3 + V4 ) . 16R

Answer:

Vout = − (4.61)

The voltage |Vout | is a maximum when [V1 V2 V3 V4 ] = [1111], in which case Vout = −

15 Rf . 16 R

To insure that |Vout | does not exceed |Vcc | = 10 V as well as to provide a safety margin of 2 V it is necessary that 8≥

15 Rf , 16 2k

which gives Rf ≤ 17.1 k. Concept Question 4-18: In a digital-to-analog converter,

what dictates the maximum value that Rf can assume? (See ) Concept Question 4-19: What is the advantage of the R–2R ladder (Fig. 4-26) over the traditional DAC (Fig. 4-25)? (See )

Rf (4V1 + 2V2 + V3 ) = −(4V1 + 2V2 + V3 ). 8R

For [V1V2V3] = [111], Vout = −7 V, whose magnitude is smaller than Vcc = 10 V. (See )

4-11 The MOSFET as a Voltage-Controlled Current Source In earlier sections, we demonstrated how op amps can be used to build buffers and amplifiers. We now examine how to realize the same outcome using MOSFETs. The simplest model of a MOSFET, which stands for metal-oxide semiconductor fieldeffect transistor, is shown in Fig. 4-27(a). The vast majority of commercial computer processors are built with MOSFETs; as mentioned in Technology Brief 1 on nanotechnology, a 2010 Intel Core processor contains over 1 billion independent MOSFETs. A MOSFET has three terminals: the gate (G), the source (S), and the drain (D). Actually, it has a fourth terminal, namely its body (B), but we will ignore it for now because for many applications it is simply connected to the ground terminal. The circuit symbol for the MOSFET may look somewhat unusual, but it is actually a stylized depiction of the physical cross section of a real MOSFET. In a real MOSFET, the gate

“book” — 2015/5/4 — 7:10 — page 220 — #38

220

CHAPTER 4

consists of a very thin layer (< 500 nm thick) of a conducting material adjacent to an even thinner layer (< 100 nm) of insulator. The insulator in turn is placed directly on the surface of a relatively large slab of semiconductor material, usually referred to as “the chip” in everyday conversation (usually made of silicon 0.5 to 1.5 mm thick). The drain and the source sections are fabricated into this semiconductor chip on either side of the gate.

 Because the gate G is separated from the rest of the transistor by the thin insulating layer, no dc current can flow from G to either D or S. 

Nonetheless, it turns out that the voltage difference between terminals G and S is key to the operation of the MOSFET. Using terminal S as a reference in Fig. 4-27(b), we denote VDS and VGS as the voltages at terminals D and G, respectively. We also denote the current that flows through the MOSFET from D to S as IDS . This simplification is justified by the assumption that no current flows through the gate node to either the drain or source node. The operation of the MOSFET can be analyzed by placing it in the simple circuit shown in Fig. 4-28(a), in which VDD is a dc power supply voltage usually set at a level close to but not greater than, the maximum rated value of VDS for the specific MOSFET model under consideration. The resistance RD is external to the MOSFET, and its role will be discussed later. The input voltage is synonymous with VGS and the output voltage is synonymous with VDS ,

Vin = VGS ,

and

Vout = VDS .

(4.62)

approximately proportional to VGS . These observations allow us to characterize the MOSFET in terms of the simple, equivalent circuit model shown in Fig. 4-28(c), which consists of a single dependent current source given by IDS = gVGS ,

4-11.1

Digital Inverter

We now will use the model given by Eq. (4.64) to demonstrate how the MOSFET can function as a digital inverter by generating an output state of “0” when the input state is “1,” and vice versa. Combining Eqs. (4.62) to (4.64) gives Vout = VDD − gRD Vin .

Since current cannot flow from G to either D or S, the only current that can flow through the MOSFET is IDS . The dependence of IDS on VGS and VDS is shown for a typical MOSFET in Fig. 4-28(b) in the form of characteristic curves displaying the response of IDS to VDS at specific values of VGS . We observe that if VDS is greater than a certain saturation threshold value VSAT , the curves assume approximately constant levels, and that these levels are

(4.65)

The constant g is a MOSFET parameter, so if we choose RD such that gRD ≈ 1, Eq. (4.65) simplifies to Vin Vout =1− . VDD VDD

(4.63)

(4.64)

where g is a MOSFET gain constant. The characteristic curves associated with this model, which is valid only if VDS exceeds VSAT , are shown in Fig. 4-28(d). Even though this equivalent circuit is very simple and more sophisticated models usually are required, it nevertheless serves as a useful approximate model for introducing some common uses of MOSFETs. In real MOSFETs, the relationship between IDS and VGS at saturation is not strictly linear. How linear the relationship is depends (in part) on the size of the transistor. Modern sub-micron transistors used in digital processors exhibit a linear relationship between IDS and VGS at saturation, whereas larger MOSFETs used for power switching may behave nonlinearly. For our purposes, the simplification denoted by Eq. (4.64) will suffice.

Moreover Vout is related to VDD by

Vout = VDD − IDS RD .

OPERATIONAL AMPLIFIERS

(4.66)

In a digital inverter, we are interested in output responses to only two input states. According to Eq. (4.66): If

Vin = 1, VDD

Vout = 0, VDD

(4.67a)

if

Vin = 0, VDD

Vout = 1. VDD

(4.67b)

and

Hence, the MOSFET circuit in Fig. 4-28(a) behaves like a digital inverter, provided the model given by Eq. (4.64) holds true and requiring that VDS exceeds VSAT . In a real circuit,

“book” — 2015/5/4 — 7:10 — page 221 — #39

4-11 THE MOSFET AS A VOLTAGE-CONTROLLED CURRENT SOURCE

VDD

D

G

IDS VDS

+

Vin _

IDS

IDS

RD

VGS

VGS = 3 V

6 mA +

5 mA

Vout

3 mA

_

1 mA

VGS = 2 V

4 mA

VGS = 1 V

2 mA

VGS very small VSAT

S Inverter circuit

(a)

IDS

RD

Vin = 3 V

3g

D

+ IDS = gVGS

+

Vin = 2 V

2g

Vout = VDS

Vin = 1 V

1g

Vin = VGS _

_

VDS

Typical characteristic curves

(b)

VDD

G

221

Vout

S (c)

Equivalent circuit

(d)

Characteristic lines of equivalent circuit

Figure 4-28: MOSFET (a) circuit, (b) characteristic curves, (c) equivalent circuit, and (d) associated characteristic lines.

Vin and Vout are not given by the simple results indicated by Eq. (4.67), but each can be categorized easily into high and low voltage values to satisfy the functionality of a digital inverter.

The NMOS inverter circuit of Fig. 4-28(a) provides the correct functionality required from a digital inverter, but it suffers from a serious power-dissipation problem. Let us consider the power consumed by RD under realistic conditions:

4-11.2

Input State 0:

NMOS versus PMOS Transistors

The MOSFET circuit of Fig. 4-28(a) actually is called an nchannel MOSFET or NMOS for short. Its operation is limited to the first quadrant in Fig. 4-28(d), where both IDS and VDS can assume positive values only. A second type of MOSFET called PMOS (p-channel MOSFET) is designed and fabricated to operate in the third quadrant, corresponding to negative values for IDS and VDS , as illustrated in Fig. 4-29. To distinguish between the two types, the symbol for PMOS includes a small open circle at terminal G.

Vin =0 VDD

IDS ≈ 0

2 PRD = IDS RD ≈ 0

(4.68a)

Input State 1: Vin =1 VDD

IDS =

VDD RD

PRD =

2 VDD . (4.68b) RD

“book” — 2015/5/4 — 7:10 — page 222 — #40

222

CHAPTER 4

OPERATIONAL AMPLIFIERS

IDS D G S

3g

VGS = 3

2g

VGS = 2

g

VGS = 1

NMOS VDS VGS = −1

−g

VGS = −2

−2g

VGS = −3

−3g

D G S PMOS

Figure 4-29: Complementary characteristic curves for NMOS and PMOS.

Heat dissipation in RD is practically zero for input state 0, but 2 /R . The value of V for input state 1, it is equal to VDD D DD , which is dictated by the MOSFET specifications, is typically on the order of volts, and RD can be made very large—on the order of k or tens of k. If RD is much larger than that, IDS becomes too small for the MOSFET to function as an inverter. For VDD on the order of 1 V and RD on the order of 10 k, PRD for an individual NMOS is on the order of 100 μW. This amount of heat generation is trivial for a single transistor, but when we consider that a typical computer processor contains on the order of 109 transistors, all confined to a relatively small volume of space, the total amount of heat that would be generated by such an NMOS-based processor would likely burn a hole through the computer! To address this heat-dissipation problem, a new technology was introduced in the 1980s called CMOS, which stands for complementary MOS.

 CMOS has revolutionized the microprocessor industry and led to the rise of the x86 family of PC processors. 

PMOS G

VDD S D

+ Vin

_

D G NMOS

+ Vout

S

_

Figure 4-30: CMOS inverter.

CMOS is a configuration that attaches an NMOS to a PMOS at their drain terminals, as shown in Fig. 4-30. The CMOS inverter provides the same functionality as the simpler NMOS inverter, but it has the distinct advantage in that it dissipates

“book” — 2015/5/4 — 7:10 — page 223 — #41

4-11 THE MOSFET AS A VOLTAGE-CONTROLLED CURRENT SOURCE

VDD = 10 V and a drain resistance RD = 1 k. The input signal vs (t) is an ac voltage with a dc-bias given by

VDD RD

Rs

υs (t) = [500 + 40 cos 300t]

+

D G

+ υs(t) _

S

223

υout(t)

_

(a) MOSFET amplifier

Note that the amplitude of the input ac signal is several orders of magnitude smaller than that of the dc voltage VDD . Apply the MOSFET equivalent model with g = 10 A/V to obtain an expression for υout (t). Solution: Upon replacing the MOSFET with its equivalent circuit, we end up with the circuit in Fig. 4-31(b). At the input side, because no current flows through Rs , it follows that

VDD

υGS (t) = υs (t),

RD

and at the output side,

D Rs + υs(t) _

G

+

iDS = gυGS

+ υout(t)

υGS

_

(μV).

S

_

(b) Equivalent circuit Figure 4-31: MOSFET amplifier circuit for Example 4-9.

υout (t) = VDD − iDS RD = VDD − gRD υGS (t) = VDD − gRD υs (t). We observe that the output voltage consists of a constant dc component (namely VDD ) and an ac component that is directly proportional to the input signal υs (t). For the element values specified in the problem, υout (t) = 10 − 10 × 103 × (500 + 40 cos 300t) × 10−6

negligible power for both input states. The significance of the inverter is in the role it plays as a basic building block for more complicated logic circuits, such as those that perform AND and OR operations.

4-11.3

MOSFETs in Analog Circuits

In addition to their use in digital circuits, MOSFETs also can be used in analog circuits as buffers and amplifiers, as demonstrated by Examples 4-9 and 4-10. As we discussed earlier in Section 4-7, a buffer is a circuit that insulates the input voltage from variations in the load resistance. Example 4-9: MOSFET Amplifier

The circuit shown in Fig. 4-31(a) is known as a commonsource amplifier and uses a MOSFET with a dc drain voltage

= 5 − 0.4 cos 300t

V.

The 5 V dc component is simply a level shift superimposed on which is a cosinusoidal signal that is identical to the input signal but is inverted and amplified by an ac gain of 104 (from 40 μV to 0.4 V).

Example 4-10: MOSFET Buffer

The circuit in Fig. 4-32(a) consists of a real voltage source (υs , Rs ) connected directly to a load resistor RL . In contrast, the circuit in Fig. 4-32(b) uses a common-drain MOSFET circuit inbetween the source and the load to buffer (insulate) the source from the load. Let us define the source as being buffered from the load if the output voltage across the load is equal to at

“book” — 2015/5/4 — 7:10 — page 224 — #42

224

CHAPTER 4

In order for υout1 /υs ≥ 0.99, it is necessary that

Rs υs

+ _

RL ≥ 99 Rs

+

RL

Source

υout1

_

or RL ≥ 9.9 k

Load

For the circuit in Fig. 4-32(c), in which the MOSFET has been replaced with its equivalent circuit, KVL gives

VDD Rs

−υs + υGS + υout2 = 0.

D

G

Also,

+ υs _

S RL

MOSFET buffer Source

+

υout2

_

Load (b) Buffer circuit

iDS = gυGS

υGS

+ _

= gRL υGS . Simultaneous solution of the two equations gives   gRL υs . υout2 = 1 + gRL

RL ≥ 9.9 ,

D G

υout2 = IDS RL

With g = 10 A/V and in order for υout2 to be no less than 0.99υs , it is necessary that

VDD Rs

(for Rs = 100 ).

(b) With MOSFET Buffer

(a) Source connected to load directly

υs

OPERATIONAL AMPLIFIERS

which is three orders of magnitude smaller than the requirement for the unbuffered circuit.

S RL

+

υout2

_

(c) Equivalent circuit Figure 4-32: Buffer circuit for Example 4-10.

least 99 percent of υs . For each circuit, determine the condition on RL that will satisfy this criterion. Assume Rs = 100  and the MOSFET gain factor g = 10 A/V.

Concept Question 4-20: What is the major advantage of

a CMOS over an NMOS circuit as a digital inverter? (See ) Concept Question 4-21: When a MOSFET is used in a buffer circuit, υout ≈ υs, where υs is the input signal voltage. So, why is it used? (See )

Exercise 4-10: In the circuit of Example 4-9, what value of RD will give the highest possible ac gain while keeping υout (t) always positive? Answer: RD = 1.85 k. (See

Solution:

C3

)

(a) No-Buffer Circuit

Exercise 4-11: Repeat Example 4-10, but require that

For the circuit in Fig. 4-32(a),

υout be at least 99.9 percent of υs . What should RL be (a) without the buffer and (b) with the buffer?

υout1 =

υ s RL . Rs + R L

Answer: (a) RL ≥ 99.9 k, (b) RL ≥ 99.9 . (See

C3

)

“book” — 2015/5/4 — 7:10 — page 225 — #43

TECHNOLOGY BRIEF 11: CIRCUIT SIMULATION SOFTWARE

225

Technology Brief 11 Circuit Simulation Software In Chapters 2 and 3 we examined all of the common methods used for analyzing linear electric circuits. In practice, these are used for designing and analyzing the many building blocks that make up larger circuits, or for obtaining approximate solutions for how more complex circuits function. In Technology Brief 1, we noted that very large scale integrated circuits (VLSI) have experienced exponential scaling for almost 50 years, so some of today’s electrical networks may include as many as 100 billion transistors! The standard circuit analysis methods available to us are accurate and applicable, but it takes a great deal of computer automation to apply them to a 100 billion–transistor network. The Multisim circuit analysis software provides an excellent start towards modeling the behavior of complex circuits. Accordingly, Multisim will be the first of two computerbased tools we will explore in this Technology Brief. Whereas Multisim is an excellent tool, it treats a circuit as a 2-D configuration, which does not account for thermal effects associated with heat generation by the circuit elements, nor possible capacitive or inductive crosscoupling of voltages between elements (through the air or insulator medium between them). To account for these effects, we need to use a sophisticated 3-D computer simulation tool. This is the subject of the second part of this Technology Brief.

Multisim Software (1) Using Simulation Tools to Calculate and Understand Engineers use electronic design automation (EDA) tools, such as Multisim, to understand the function of a circuit and calculate its response. Consider the simple example shown in Fig. TF11-1(a), and let us assume we need to determine what voltage Vr would be measured by the voltmeter shown in the circuit. In this case, because the circuit is very simple, we can analyze it by hand or we can implement it and solve it by Multisim (Fig. TF11-1(b)). But if the circuit has more than five nodes, the byhand approach becomes tedious, and the Multisim option becomes far more practical.

Vr 3.8 kΩ

+

5V_

1 kΩ

1.2 kΩ 3 kΩ

+ _5V

(a) Circuit

(b) Multisim layout Figure TF11-1: Two-source circuit and Multisim representation using switches to switch one or both voltage sources on or off.

(2) Using Simulation Tools to Lay Out a Circuit Once a circuit has been designed, we can either build it on a protoboard or, alternatively, we can have a circuit board built for it and then solder the parts to the board to create the circuit. Printed circuit board (PCB) layout tools help us plan the circuit layout and routing architecture, which often are multiple layers deep, as in the circuit of Fig. TF11-2. When using silicon chips, for example, these designs involve hundreds, millions, or trillions of components arranged in one or more layers, and carrying thousands of simultaneous signals throughout the circuit, all acting together to obtain the desired voltage and/or current output of the circuit. Classic EDA tools (such as Multisim) begin with a graphical user interface (GUI) that allows users to specify what type of circuit elements (sources, resistors, switches, etc.) are needed and how they are connected together. Circuits made up of several elements can often be grouped or bundled together and stored in libraries for later reuse. Often, libraries of complex

“book” — 2015/5/4 — 7:10 — page 226 — #44

226

TECHNOLOGY BRIEF 11: CIRCUIT SIMULATION SOFTWARE

Figure TF11-2: Multilayer PCB layout, with each layer assigned a different color. Holes and solder pads are planned for each chip and component attached to the board, and multilayer routing built into the circuit board connects them all together. (Courtesy of ZYPEX Inc.)

circuits (such as the core of a computer processor) are shared or purchased to reduce engineering design time. For circuits whose design can be expressed as either logical rules or a desired logical function—primarily digital circuits—modern software tools transform circuit design into an exercise in writing code. In essence, programs can be written in hardware description languages (HDL), which define the structure and/or operation of digital circuits. The program is then executed and a circuit description suitable for manufacture, or instantiation into a field-programmable gate array (FPGA), is synthesized. Programming in HDLs is similar to assembly language or C coding, although major

differences exist. Most modern complex digital circuits are designed, simulated, and synthesized with the aid of HDL tools. Once the elements and their connections are defined, they are then modeled with either more or less detail (by specifying tolerance levels or other relevant parameters) depending on the level of accuracy needed. Simulation results are only as good as the circuit model and input parameters, so this is a very important consideration when using EDA software. The more detailed the model, the more accurate the results can be expected to be, but also the longer it takes the simulation to run. Consider, for example, the ideal and the more realistic models

“book” — 2015/5/4 — 7:10 — page 227 — #45

TECHNOLOGY BRIEF 11: CIRCUIT SIMULATION SOFTWARE for voltage and current sources listed in Table 1-5. The realistic source models are certainly more accurate than the ideal models, but even the “realistic” models are approximate, because they neglect nonlinearities, stray capacitance and inductance, and potential feedback loops within the sources. For many applications, the ideal model is sufficient, for others the first-order realistic model (including a resistor) is sufficient, but for others, a more detailed nonlinear model is required. How do you, the engineer, know what model to use? The intuition and knowledge gained from working with the common circuit analysis tools from Chapters 2 and 3 help you determine when you may or may not need a more realistic model. Often, we will first try a simplified model, and then one that is slightly more realistic. If there is minimal change, we do not go on to a more complex model, but if there is substantial change, we may try more and more realistic models (each requiring more time and memory for the software to run), until the result converges and we are satisfied that we have modeled the real system at hand. Now let’s consider VLSI circuits involving trillions of transistors. Even with relatively simple models of the transistor (such as the BJT in Section 3-9 or MOSFET in Section 4-11), there are still more unknowns than we generally care to wait for the computer to solve. In this case, two simplifications are essential. First, we must break the circuit down into functional blocks, so we can design each block individually and cascade or connect the blocks together.We have already seen simple examples of doing this using the Thevenin ´ equivalent circuit technique. Thevenin ´ is also used this way in much larger circuits, including VLSI designs. Second, we must simplify the models we use for each circuit element. Fortunately (or perhaps necessarily!) the largest circuits electrical engineers design are digital circuits, for which we can use the simplest models of all. We can assume that all voltages are either high/on (digital 1) or low/off (digital 0). This flexibility in the voltages allows us to use much simpler models. The transistor, for example, can be modeled as just a switch (on or off), or just as a resistor that is switched in or out of the circuit. Assuming all voltages are either on or off is the simplest assumption. We also can model them as on/off or in transition between on and off. The transition (which is actually a bouncy switch) can be modeled as a linear slope from low to high or high to low. The length of this slope is the rise time of the transition, and the faster the rise time, the faster the circuit can send data.

227

3-D Modeling Tools Model-based EDA tools define how a circuit is supposed to function electrically, but sometimes effects not included in the models come into play to make the circuit malfunction.Two of these that are particularly relevant are associated with thermal problems and coupling problems. We know that resistors and other devices are designed with specific power ratings. The power rating is related to the size and material the resistors are made of and their ability to withstand the heat generated by current moving through them. If we start pushing all of the elements of the circuit to their maximum capability, their interactions (hot chips next to other hot chips) may make the most vulnerable of these parts fail. But how do we determine which parts are the most vulnerable, and what solution can we offer to mitigate the heat problem? 3-D simulation tools help us to identify these potential problems or (all too often) diagnose them when they occur. The 3-D simulation process starts with the physical model of a given part, such as the high-speed IC package shown in Fig. TF11-3(a). The spatial distributions of electrical voltage and current are then modeled for part or all of the package, as shown in Fig. TF11-3(b). The current density at a given location is representative of what the temperature will be at that location. If overheating were to occur, it would most likely occur at the points with the highest current. More detailed thermal modeling can include the effects of heat sinks, fans, and other cooling effects. The voltage is used to calculate coupling between nearby electrical signals (such as two adjacent legs of this package). Another interesting circuit simulation is shown in Fig. TF11-4, which displays the amount of power radiated by a crescent antenna.

So WHY Should You Learn the Circuit Analysis Methods Introduced in This Book? Having learned how to apply the various circuit analysis tools covered in this book thus far, you may wonder why you need to learn so many different methods when they all can give you the same result. And now that you have read this Technology Brief and seen that you can use a computer to analyze circuits, you may wonder why you need to learn these analytical methods at all! While it is true that automated tools are essential for testing circuits used in practical applications, it is equally true that the success of the design process is highly coupled to one’s understanding of the fundamental

“book” — 2015/5/4 — 7:10 — page 228 — #46

228

TECHNOLOGY BRIEF 11: CIRCUIT SIMULATION SOFTWARE

(a) Physical package

(b) Current density contour

Figure TF11-3: High-speed IC package and contour and vector plot of the current density flowing through it at 5 GHz. The brighter/redder colors show higher current density (A/m2 ) (which also results in higher temperature) than the darker/bluer colors. The arrows show the direction in which the current is flowing, and the size of the arrow is also proportional to the r IC Package Simulation.) magnitude of the current density. (Courtesy: CST MICROWAVE STUDIO

concepts in circuit analysis and design. Designing a new circuit to address a specified application is a creative endeavor that relies on one’s past experience and fluency in circuit behavior and performance. Once an initial circuit

configuration has been developed, computer simulation tools are then used to fine-tune the design and optimize the circuit performance.

Figure TF11-4: This 3D electromagnetic simulation was used to evaluate the fields (in this case the square of the electric field, which is proportional to power) in the nanocrescent antenna shown in Technology Brief 1. We can see the strong fields at the tips (because charge congregates there), and also in the center. (Credit: Miguel Rodriguez.)

“book” — 2015/5/4 — 7:10 — page 229 — #47

4-12 APPLICATION NOTE: NEURAL PROBES

229 aspects of brain development and operation, but they also are beginning to see use in clinical applications for the treatment of chronic neurological disorders, such as Parkinson’s disease (see Technology Briefs 17 and 32 on neural stimulation and computer-brain interfaces, respectively). Because these voltage signals are so small, on-board amplification, noise-removal, and analog-to-digital circuitry are needed to process the signal from the brain to the recording device. Example 4-11: Neural Probe

Figure 4-33: Three-dimensional neural probe (5 mm × 5 mm × 3 mm). (Courtesy of Prof. Ken Wise and Gayatri Perlin, University of Michigan.)

4-12 Application Note: Neural Probes The human brain is composed, in part, of interconnected networks of individual, information-processing cells known as neurons. There are about one trillion (1012 ) neurons in the human brain with each neuron having on average 7000 connections to other neurons. Although the working of the neural system is well beyond the scope of this book, it is important to note that when a neuron transmits information, it causes a change in the concentrations of various ions in its vicinity. This movement of ions gives rise to an electric current through the neuron’s membrane which in turn generates a change in potential (voltage) between various parts of the cell and its surroundings. Thus, when a given neuron fires, a small (∼ 100 mV) but detectable potential drop develops between the cell and its surroundings. Over the past few decades, various types of devices were built for measuring this electrical phenomenon in neurons. In recent years, however, the field has achieved phenomenal success due in part to the successful development of neural probes (also known as neural interfaces) with very high sensitivities. An example of a 3-dimensional probe is shown in Fig. 4-33. It consists of a 2-D array of very thin probes—each instrumented with a sensor at each of several locations along its length. With such a probe, it is now possible to measure the action potentials of firing neurons at a large number of brain locations simultaneously. Modern neural interface systems also have been developed to stimulate or change the electrical state of specific neurons, thereby affecting their operation in the brain. These types of devices not only offer the potential of unraveling

The neural probe shown in Fig. 4-34 consists of a long shank at the end of which lie two metal electrodes. This shank is inserted a short distance into the brain and the signal coming from these electrodes is recorded. For simplicity, we will model the brain activity between the two probes just like a realistic voltage source Vs in series with a resistance Rs . The source produces inverted pulses with −100 mV amplitudes. Note that neither Va nor Vb are grounded relative to the ground level of the circuit. The neural signal needs to be inverted and amplified so that it can be presented to an analog-to-digital converter (ADC) which only operates in the 0 to 5 V range. Design the amplifier circuit. Solution: The input signal is represented by the difference between Va and Vb , and since neither of those terminals is grounded, some sort of differential amplifier is the logical choice for the intended application. The amplifier should invert the input signal and amplify it into the 0 to 5 V range required by the ADC. Given these constraints, we propose to use the op-amp instrumentation amplifier circuit of Fig. 4-23 with Va as input υ1 and Vb as input υ2 . The amplifier output is proportional to (υ2 −υ1 ), so the choice of connections we made will realize the inversion requirement automatically. According to Eq. (4.56), if we choose the circuit resistors such that R1 = R3 = R4 = R5 = R, the output voltage is given by   2R (υ2 − υ1 ) υo = 1 + R2     2R 2R = 1+ (Vb − Va ) = − 1 + (Va − Vb ). R2 R2 To amplify (Va − Vb ) from −100 mV to +5 V, the ratio (R/R2 ) should be chosen such that   2R × (−100 × 10−3 ) 5=− 1+ R2 or, equivalently, R = 24.5. R2 If we set R = 100 k, then R2 should be 4.08 k. This will yield a 5 V pulse to the ADC every time a −100 mV pulse is generated by the neuron.

“book” — 2015/5/4 — 7:10 — page 230 — #48

230

CHAPTER 4

OPERATIONAL AMPLIFIERS

Probe

Va Amplifier

ADC

To computer

Vb Probe

Neuron

Va − Vb 1 ms t Rs Vs

Va

Vb

−100 mV

Figure 4-34: Neural-probe circuit for Example 4-11.

4-13 Multisim Analysis One of the most attractive features of Multisim is its interactivesimulation mode, which we began to utilize in Sections 2-7 and 3-8. The simulation mode allows you to connect virtual test instruments to your circuit and to operate them in real time as Multisim simulates the circuit behavior. In this section, we will explore this feature with an op-amp circuit and two MOSFET circuits.

4-13.1

Op Amps and Virtual Instruments

The circuit shown in Fig. 4-35 uses a resistive Wheatstone bridge (Section 2-5) to detect the change of resistance induced in a sensor modeled as a variable resistor (see Technology Brief 4 on resistive sensors). The output of the bridge is fed into a pair of voltage followers and then into a differential amplifier. The circuit can be constructed and tested in Multisim using the components listed in Table 4-5. The resistance value of the potentiometer component is adjustable with a keystroke

(the default is the key “a” to change the resistance in one direction and the default key combination Shift-a to change the resistance in the opposite direction) or by using the mouse slider under the component. In order to observe how changes in the potentiometer cause changes in the output, we need to connect the output to an oscilloscope. Multisim provides several oscilloscopes to choose from, including a generic instrument and virtual versions of commercial oscilloscopes made by Agilent and Tektronix. For starters, it is easiest to use the generic instrument by selecting Simulate → Instruments → Oscilloscope, or by selecting and dragging an oscilloscope from the instrument dock. Figure 4-36 shows the complete circuit drawn in Multisim. The power supplies for the op amps can be found under Components → Sources → POWER SOURCES → VDD (or VSS). Once placed, double-click the VDD (or VSS) component, select the values tab and set the voltage to 15 V for VDD and −15 V for VSS. Once the circuit is complete, you can begin the simulation by pressing F5 (or Simulate → Run) and pause it by pressing F6. Double-click on the oscilloscope element in the schematic to bring up the

“book” — 2015/5/4 — 7:10 — page 231 — #49

4-13

MULTISIM ANALYSIS

231

R R

Sensor + + _ −

~ R

15 kΩ R

1V R

R R = 1.5 kΩ

Range: 0 to 3 kΩ

+ _ + _

+ _

R

Vout 15 kΩ

Sensor bridge

Voltage follower (Gain = 1)

Differential amplifier (Gain = 10)

Figure 4-35: Wheatstone-bridge op-amp circuit. oscilloscope’s screen. The output voltage should be visible as Channel A in the oscilloscope window. In order to get a good view of the trace, you might need to adjust both its timebase and voltage scale using the controls found at the bottom of the Oscilloscope window. Observe the change in the amplitude of the output by shifting the resistance value of the sensor potentiometer. With Multisim, you can modify different parts of the circuit and observe the consequent changes in behavior. Make sure to stop your simulation (not just pause it) before changing components or wiring. Concept Question 4-22: What types of Multisim

instruments are available for testing a circuit? (See

)

Concept Question 4-23: Explain what the timebase is on

the oscilloscope. (See

)

Exercise 4-12: Why are the voltage followers necessary in

the circuit of Fig. 4-36? Remove them from the Multisim circuit and connect the resistive bridge directly to the two inputs of the differential amplifier. How does the output vary with the potentiometer setting? Answer: (See

)

4-13.2 The Digital Inverter The MOSFET inverter introduced in Section 4-11.2 provides a good opportunity to explore the difference between steadystate and time-dependent analysis techniques. Consider again the MOSFET digital inverter of Fig. 4-30. When analyzing this type of logic gate, we usually are interested in both the response of the output voltage to a change in input voltage and in how fast the gate generates the output voltage in response to a change in input voltage. Both types of analyses are possible with Multisim.

Table 4-5: List of Multisim components for the circuit in Fig. 4-35. Component

Group

Family

Quantity

Description

1.5 k

Basic

Resistor

7

1.5 k resistor

15 k

Basic

Resistor

2

15 k resistor

3k

Basic

Variable resistor

1

3 k resistor

OP AMP 5T VIRTUAL

Analog

Analog Virtual

3

Ideal op amp with 5 terminals

AC POWER

Sources

Power Sources

1

1 V ac source, 60 Hz

VDD

Sources

Power Sources

1

15 V supply

VSS

Sources

Power Sources

1

−15 V supply

“book” — 2015/5/4 — 7:10 — page 232 — #50

232

CHAPTER 4

OPERATIONAL AMPLIFIERS

Figure 4-36: Multisim window of the circuit of Fig. 4-35. The oscilloscope trace shows the 60 Hz waveform of the output voltage. Had the voltage source been a dc source, the oscilloscope trace would have been a horizontal line.

Figure 4-37 shows a MOSFET inverter circuit in Multisim. To draw this circuit, you need the components listed in Table 4-6. Transient Analysis We can use a function generator (Simulate → Instruments → Function Generator) to observe the inverter output as a function of time. Double-click on the function generator to bring up its control window. Set the function generator to Square Wave mode with a frequency of 1 kHz, amplitude of 2.5 V,

and an offset of 1.25 V. This will generate a 0–2.5 V squarewave input. The input and output can be plotted separately as a function of time using Simulate → Analyses → Transient Analysis. Whereas in Interactive Simulation the course of time is open ended (by default it is limited to a duration of 1×1030 s), when using Transient Analysis we can define the start and stop times. Maintain the start time at 0 s, set the final time to 0.005 s, and under the Output tab select the input voltage V(1) as the voltage to plot. Click Simulate. The input voltage is plotted as a function of time, as in Fig. 4-38(a). Repeat the simulation after removing V(1) and adding V(2) under the Output tab.

“book” — 2015/5/4 — 7:10 — page 233 — #51

4-13

MULTISIM ANALYSIS

233

Table 4-6: Components for the circuit in Fig. 4-37. Component

Group

Family

MOS N

Transistors

Transistors VIRTUAL

Quantity 1

Description 3-terminal N-MOSFET

MOS P

Transistors

Transistors VIRTUAL

1

3-terminal P-MOSFET

VDD

Sources

Power Sources

1

2.5 V supply

GND

Sources

Power Sources

2

Ground node

(a)

Input voltage

(b)

Output voltage

Figure 4-38: Input and output voltages V(1) and V(2) in the Figure 4-37: Multisim equivalent of the MOSFET circuit of

circuit of Fig. 4-37 as a function of time.

Fig. 4-30.

Figure 4-38(b) shows the output voltage as a function of time. The input and output plots are essentially mirror images of one another.

Steady-State Analysis In order to analyze the steady-state output behavior, we first must remove the function generator and replace it with a

“book” — 2015/5/4 — 7:10 — page 234 — #52

234

CHAPTER 4

DC Sweep

OPERATIONAL AMPLIFIERS

Exercise 4-13: The IV Analyzer is another useful Multisim instrument for analyzing circuit performance. To demonstrate its utility, let us use it to generate characteristic curves for an NMOS transistor similar to those in Fig. 4-28(b). Figure E4.13(a) shows an NMOS connected to an IV Analyzer. The instrument sweeps through a range of gate (G) voltages and generates a current-versus-voltage (IV) plot between the drain (D) and source (S) for each gate voltage. Show that the display of the IV analyzer is the same as that shown in Fig. E4.13(b). Answer: (See

C3

)

Figure 4-39: Output response of the MOSFET inverter circuit of Fig. 4-37 as a function of the amplitude of the input voltage.

dc voltage source. The actual voltage value of the source is unimportant. Once wired, select Simulate → Analyses → DC Sweep. This analysis is similar to the DC Operating Point Analysis, but it sweeps through a range of voltages at a node of your choice and solves for the resultant steady-state voltage (or current) at any other node you select. In this way, you can generate and plot input-output relationships for circuits and components. Choose the source name vv1 as the input and enter 0 V, 2.5 V, and 0.01 V for the start, stop, and increment values, respectively. Under the Output tab, select the output voltage V(2) as the voltage to plot. Click Simulate. Figure 4-39 shows that the output displays the expected inverter behavior: an input in the 0 to 500 mV range generates an output of ∼ 2.5 V; conversely, when the input is in the range between 2 and 2.5 V, the circuit generates an output voltage of ∼ 0 V. In between, we see a gradual transition zone. Concept Question 4-24: How do the DC Operating

Point Analysis, Transient Analysis, and DC Sweep analyses differ? (See ) Concept Question 4-25: How many types of waveforms can the generic function-generator instrument provide? (See )

(a)

VGS = 5 V VGS = 3.75 V VGS = 2.5 V VGS = 1.25 V VGS = 0

(b) Figure E4.13 (a) Circuit schematic and (b) IV analyzer

traces for IDS versus VDS at selected values of VGS .

“book” — 2015/5/4 — 7:10 — page 235 — #53

4-13

MULTISIM ANALYSIS

235

Summary Concepts • Despite its complex circuit architecture, the op amp can be modeled in terms of a relatively simple, linear equivalent circuit. • The ideal op amp has infinite gain A, infinite input resistance Ri , and zero output resistance Ro . • Through resistive feedback connections between its output and its two inputs, the op amp can be made to amplify, sum, and subtract multiple input signals. • Multistage op-amp circuits can be configured to support a variety of signal-processing functions.

• Cascaded circuit blocks can be analyzed or designed individually and then combined together if Ro of the first circuit is much smaller than Ri of the second circuit. • Buffers are used to increase Ri of the followup circuit. • The instrumentation amplifier is a high-gain, highsensitivity detector of small signals, making it particularly suitable for sensing deviations from reference conditions. • Multisim can accommodate op-amp circuits and simulate their input-output responses.

Mathematical and Physical Models υp = υn ip = in = 0

Ideal op amp

Noninverting

amp∗

Inverting amp∗

Summing

amp∗

Important Terms action potential ADC adder bit buffer circuit gain closed-loop gain CMOS complementary MOS current constraint difference amplifier digital inverter digital-to-analog converter DIP configuration drain dynamic range feedback feedback resistance gain-control resistance

υo R1 + R2 G= = υs R2   υo Rf G= =− υs Rs   υ1 υ2 υo = −Rf + R1 R2

Difference amp∗

υo = G2 υ2 + G1 υ2

Voltage follower∗

υo = υs 

2R Instrumentation amp υo = 1 + R2 (with gain-control resistor R2 )

(υ2 − υ1 )

Vout = VDD − gRD Vin

MOSFET ∗ See Table



4-3.

Provide definitions or explain the meaning of the following terms: gate ideal op-amp current constraint ideal op-amp voltage constraint input resistance input source resistance instrumentation amplifier inverter inverting inverting adder inverting amplifier inverting input inverting summing amplifier IV Analyzer least significant bit linear linear dynamic range loading

metal-oxide semiconductor field-effect transistor MOSFET MOSFET gain constant most significant bit negative feedback negative saturation neural interface neural probe neuron NMOS noninverting amplifier noninverting noninverting input noninverting summing amplifier oscilloscope op amp op-amp gain open-loop gain

operational amplifier output resistance overloading percent clipping PMOS positive feedback positive saturation R–2R ladder saturation threshold value scaled inverting adder sensor signal-processing circuit source subtraction summing amplifier unity gain amplifier voltage constraint voltage follower voltage rails

“book” — 2015/5/4 — 7:10 — page 236 — #54

236

CHAPTER 4

PROBLEMS

(a) Use the equivalent-circuit model of Fig. 4-6 to obtain an expression for the closed-loop gain G = υo /υs in terms of Rs , Ri , Ro , RL , Rf , and A.

Sections 4-1 and 4-2: Op-Amp Characteristics and Negative Feedback

(b) Determine the value of G for Rs = 10 , Ri = 10 M, Rf = 1 k, Ro = 50 , RL = 1 k, and A = 106 .

*4.1 An op amp with an open-loop gain of 106 and Vcc = 12 V has an inverting-input voltage of 20 μV and a noninvertinginput voltage of 10 μV. What is its output voltage?

(c) Simplify the expression for G obtained in (a) by letting A → ∞, Ri → ∞, and Ro → 0 (ideal op-amp model). *(d) Evaluate the approximate expression obtained in (c) and compare the result with the value obtained in (b).

4.2 An op amp with an open-loop gain of 6 × 105 and Vcc = 10 V has an output voltage of 3 V. If the voltage at the inverting input is −1 μV, what is the magnitude of the noninverting-input voltage?

υp

Rs

(b) Simplify the expression by applying the ideal op-amp model (taking A → ∞, Ri → ∞, and Ro → 0).

is

+ _

Rs

Figure P4.5: Circuit for Problem 4.5.

4.6 The inverting-amplifier circuit shown in Fig. P4.6 uses a resistor Rf to provide feedback from the output terminal to the inverting-input terminal. Answer(s) available in Appendix G.

(a) Use the op-amp equivalent-circuit model to develop an expression for G = υo /υs . (b) Simplify the expression by applying the ideal op-amp model parameters, namely A → ∞, Ri → ∞, and Ro → 0.

_

RL



RL

For the circuit in Fig. P4.7:

R1 iL

Rf

Figure P4.6: Circuit for Problem 4.6.

4.7

(a) Use the model given in Fig. 4-6 to develop an expression for the current gain Gi = iL / is .

υo

+ υs _

4.4 With its noninverting-input voltage at 10 μV, the output voltage of an op amp is −15 V. If A = 5 × 105 and Vcc = 15 V, can you determine the magnitude of the inverting-input voltage? If not, can you determine its possible range? 4.5 For the op-amp circuit shown in Fig. P4.5:

+ _

υn

*4.3 What is the output voltage for an op amp whose noninverting input is connected to ground and its invertinginput voltage is 4 mV? Assume that the op-amp open-loop gain is 2 × 105 and its supply voltage is Vcc = 10 V.

υp υn

OPERATIONAL AMPLIFIERS

+ υs _

+

υo RL

Figure P4.7: Circuit for Problem 4.7.

4.8 The op-amp circuit shown in Fig. P4.8 has a constant dc voltage of 6 V at the noninverting input. The inverting input is the sum of two voltage sources consisting of a 6 V dc source and a small time-varying signal υs .

“book” — 2015/5/4 — 7:10 — page 237 — #55

PROBLEMS

237

(a) Use the op-amp equivalent-circuit model given in Fig. 4-6 to develop an expression for υo .

Rf a

(b) Simplify the expression by applying the ideal op-amp model, which lets A → ∞, Ri → ∞, and Ro → 0.

υs

+ _

υo

R3

RL

+ _

_

R2 _

+ 6V _

+ 6V _

R1 +

υs

R4

b

Figure P4.10: Circuit for Problem 4.10.

4.11 Determine the output voltage for the circuit in Fig. P4.11 and specify the linear range for υs , given that Vcc = 15 V and V0 = 0.

Figure P4.8: Circuit for Problem 4.8.

200 kΩ

Sections 4-3 and 4-4: Ideal Op Amp and Inverting Amp

_

2 kΩ Assume all op amps to be ideal from here on forward.

+ 100 kΩ

*4.9 The supply voltage of the op amp in the circuit of Fig. P4.9 is 16 V. If RL = 3 k, assign a resistance value to Rf so that the circuit would deliver 75 mW of power to RL .

50 Ω

+ 3V _

+ _

υs

Rf

υo Vcc = 15 V

+ _

V0

Figure P4.11: Circuit for Problems 4.11 and 4.12.

Vcc = 16 V 4.12

4 kΩ

υo

+

RL

Repeat Problem 4.11 for V0 = 0.1 V.

*4.13 Obtain an expression for the voltage gain G = υo /υs for the circuit in Fig. P4.13.

R2

Figure P4.9: Circuit for Problem 4.9.

_ 4.10 In the circuit of Fig. P4.10, a bridge circuit is connected at the input side of an inverting op-amp circuit. (a) Obtain the Th´evenin equivalent at terminals (a, b) for the bridge circuit.

R1

R3

+ Rs

υo RL

υs

Figure P4.13: Circuit for Problem 4.13.

(b) Use the result in (a) to obtain an expression for G = υo /υs . (c) Evaluate G for R1 = R4 = 100 , R2 = R3 = 101 , and Rf = 100 k.

4.14 For the op-amp circuit shown in Fig. P4.14: (a) Obtain an expression for the current gain Gi = iL / is .

“book” — 2015/5/4 — 7:10 — page 238 — #56

238

CHAPTER 4

OPERATIONAL AMPLIFIERS

*(b) If RL = 12 k, choose Rf so that Gi = −15.

Rf

_

Rf 600 Ω

_ is

iL

+

Rs

Vcc = 7 V

400 Ω

RL

υo

+

1200 Ω

1200 Ω

+ υs _

Figure P4.14: Circuit for Problem 4.14. Figure P4.18: Circuit for Problems 4.18 and 4.19.

+ _ 5 kΩ

Vcc = 6 V υo 4 kΩ 70 kΩ

20 kΩ

υs

υL

Repeat Problem 4.18 for Rf = 0.

*4.19

4.20 Determine the linear range of the source υs in the circuit of Fig. P4.20.

1.2 kΩ

RL

_

10 kΩ 200 Ω

Figure P4.15: Circuit for Problems 4.15 and 4.16.

υs

*4.16 For the circuit of Fig. P4.15, what should the resistance value of RL be so as to have maximum transfer of power into it? 4.17 Determine υo across the 10 k resistor in the circuit of Fig. P4.17.

2 kΩ 1V

_ +

_ +

50 Ω 5V

+ _

υo

υo Vcc = 12 V

*4.21 Repeat Problem 4.20 after replacing the 2 V dc source in Fig. P4.20 with a short circuit. 4.22 The circuit in Fig. P4.22 uses a potentiometer whose total resistance is R = 10 k with the upper section being βR and the bottom section (1 − β)R. The stylus can change β from 0 to 0.9. Obtain an expression for G = υo /υs in terms of β and evaluate the range of G (as β is varied over its own allowable range).

υs

100 Ω

+ _

βR

(1 − β)R

υo 678

4.18 Evaluate G = υo /υs for the circuit in Fig. P4.18, and specify the linear range of υs . Assume Rf = 2400 .

+ _ 2V

+

Figure P4.20: Circuit for Problems 4.20 and 4.21.

10 kΩ

Figure P4.17: Circuit for Problem 4.17.

400 Ω

876 876

4.15 Determine the gain G = υL /υs for the circuit in Fig. P4.15 and specify the linear range of υs for RL = 4 k.

R = 10 kΩ

Figure P4.22: Circuit for Problem 4.22.

“book” — 2015/5/4 — 7:10 — page 239 — #57

PROBLEMS

239

4.23 For the circuit in Fig. P4.23, obtain an expression for voltage gain G = υo /υs .

4.28 For the circuit in Fig. P4.28, generate a plot for υL as a function of υs over the full linear range of υs .

5 kΩ 10 kΩ

_ +

υo 20 kΩ

4 kΩ 6 kΩ υs

+ 0.5 V _

Figure P4.23: Circuit for Problem 4.23.

*4.24

_

4 kΩ

Find the value of υo in the circuit in Fig. P4.24.

4V + _

υs

+ _

4.27 Design an op-amp circuit that performs an averaging operation of five inputs υ1 to υ5 .

+

υL RL

Vcc = 12 V

Figure P4.28: Circuit for Problem 4.28.

4 kΩ 4.29 Relate υo in the circuit of Fig. P4.29 to υs and specify the linear range of υs . Assume V0 = 0.

6 kΩ 6 kΩ 2 mA

_ +

υo

2 kΩ 5V

8 kΩ

+ _

Vcc = 16 V

_ 2 kΩ

Figure P4.24: Circuit for Problem 4.24.

υs

4.25 Determine the linear range of υs for the circuit in Fig. P4.25.

3V

8 kΩ

4 kΩ + _

4V

+ _

+

+ _ V0

io

υo RL

Figure P4.29: Circuit for Problems 4.29 through 4.31.

Vcc = 16 V 10 kΩ 2V

+ _

υs

+ _

+ _

υo 15 kΩ

20 kΩ

Figure P4.25: Circuit for Problem 4.25.

*4.30

Repeat Problem 4.29 for V0 = 6 V.

4.31 Determine the current io flowing into the op-amp of the circuit in Fig. P4.29 under the conditions υs = 0.5 V, V0 = 0, and RL = 10 k. 4.32 Design a circuit containing a single op amp that can perform the operation υo = 3 × 104 (i2 − i1 ), where i2 and i1 are input current sources.

Sections 4-5 and 4-6: Summing and Difference Amplifiers 4.26 If R2 = 4 k, select values for Rs1 , Rs2 , and R1 in the circuit of Fig. 4-15 so that υo = 3υ1 + 5υ2 .

4.33 Design a circuit that can perform the operation υo = 3υ1 + 4υ2 − 5υ3 − 8υ4 , where υ1 to υ4 are input voltage signals.

“book” — 2015/5/4 — 7:10 — page 240 — #58

240

CHAPTER 4 Relate υo in the circuit of Fig. P4.34 to υ1 , υ2 , and υ3 .

4.34

OPERATIONAL AMPLIFIERS

Rf

4 kΩ

R5 R4

R1 υ1

Vcc = 10 V

_ R2 υ2

R3

3 kΩ

_ +

υo

+

RL 7V

υ3

+ _

+ _

+ _

6V

υo

4V

Figure P4.34: Circuit for Problem 4.34. Figure P4.37: Circuit for Problem 4.37.

*4.35 For the circuit in Fig. P4.35, obtain an expression for υo in terms of υ1 , υ2 , and the four resistors. Evaluate υo if υ1 = 0.1 V, υ2 = 0.5 V, R1 = 100 , R2 = 200 , R3 = 2.4 k, and R4 = 1.2 k.

*4.38 Determine υo and the power dissipated in RL in the circuit of Fig. P4.38.

5 kΩ 2V _

R3 υ1 υ2

_ R4

R2

7 kΩ

+

R1

Vcc = 16 V υo

+

4V

+ _

_ +

3 kΩ

3 kΩ

4 kΩ

υo

2 kΩ 2 kΩ

RL

Figure P4.35: Circuit for Problem 4.35. Figure P4.38: Circuit for Problem 4.38.

4.36

Find the value of υo in the circuit in Fig. P4.36.

5 kΩ 2 kΩ

3V

+ _

9V

+ _

4.39 The circuit in Fig. P4.39 contains two single-pole singlethrow switches, S1 and S2 . Determine the closed-circuit gain G = υo /υs for each of the four possible closed/open switch combinations.

4 kΩ

_ +

υo 24 kΩ

6 kΩ

_

4 kΩ

S2

6 kΩ Figure P4.36: Circuit for Problem 4.36.

4.37 Find the range of Rf for which the op amp in the circuit of Fig. P4.37 does not saturate.

S1

+

υs 6 kΩ

6 kΩ

Figure P4.39: Circuit for Problem 4.39.

υo

“book” — 2015/5/4 — 7:10 — page 241 — #59

PROBLEMS

241

Section 4-8: Op-Amp Signal-Processing Circuits

Relate υo in the circuit of Fig. P4.46 to υs .

*4.46

4.40 Develop a block-diagram representation for the circuit in Fig. P4.40 for υs2 = υs3 = 0 and

υs

*(a) R1 = open circuit

Rs

+

(b) R1 = 10 k.

24 kΩ υs1 υs2

400 kΩ

_

4 kΩ

20 kΩ

+

R1

2 kΩ

υs 3

R3

_

R2

_ +

R1

υo

+

4.47 In the circuit of Fig. P4.47, op amp 1 receives feedback at its input from its own output as well as from the output of op amp 2. Relate υo to υs .

Rf3

4.41 Develop a block-diagram representation for the circuit in Fig. P4.40 for υs3 = 0 and R1 = ∞.

Rf1 υs

Rs1

Rf2

_ Op _

4.42 Develop a block-diagram representation for the circuit in Fig. P4.40 for υs2 = 0 and R1 = ∞.

Rs2

Amp 1

+ +

(b) Specify the linear range of υs .

Figure P4.47: Circuit for Problem 4.47.

(c) Determine υo for υs = 0.3 V and RL = 10 k. 4.48

Relate υo in the circuit of Fig. P4.48 to υ1 and υ2 .

80 kΩ

+

+

_

0.4 V

2 kΩ

8 kΩ

Vcc = 12 V

_

υo

+

(a) Develop a block-diagram representation with RL as a variable parameter.

4 kΩ

_ Op Amp + 2

For the circuit in Fig. P4.43:

υs

RL

Figure P4.46: Circuit for Problem 4.46.

_

Figure P4.40: Circuit for Problems 4.40 through 4.42.

4.43

υo

Vcc = 12 V

+ _

10 kΩ

2 kΩ υo

υ1

RL

Figure P4.43: Circuit for Problem 4.43.

4.44 Design an op-amp circuit that can perform the operation υo = 12υs1 + 3υs2 , while simultaneously presenting an input resistance of 50 k on the input side for source υs1 and an input resistance of 25 k on the input side for source υs2 . 4.45 Design an op-amp circuit that can perform the operation υo = 4υs1 − 3υs2 , while simultaneously presenting an input resistance of 10 k on the input side for source υs1 and an input resistance of 5 k on the input side for source υs2 .

_ + 20 kΩ

+ _

υ2 2 kΩ

40 kΩ

_ +

0.5 kΩ 4 kΩ

_ +

10 kΩ

υo

10 kΩ 40 kΩ

+ _

0.5 kΩ

16 kΩ Figure P4.48: Circuit for Problem 4.48.

4.49 Design an op-amp circuit that can perform the operation io = (30i1 − 8i2 + 0.6) A where i1 and i2 are input current sources.

“book” — 2015/5/4 — 7:10 — page 242 — #60

242

CHAPTER 4 Relate the output voltage υo in Fig. P4.50 to υs .

10 kΩ

+ _

12 kΩ

υo 8 kΩ

6 kΩ 4 kΩ

3V _ 5V _





4Ω 4Ω

_ + 5V _

2 kΩ

+

+ _

Solve for υo in the circuit in Fig. P4.53.

+

υs 50 kΩ

4.53

+

*4.50

OPERATIONAL AMPLIFIERS



_ +

υo





+ _



Figure P4.50: Circuit for Problem 4.50.

3Ω 4.51



Solve for υo in terms of υs for the circuit in Fig. P4.51. Figure P4.53: Circuit for Problem 4.53.

10 kΩ 2 kΩ

+ _

υs

4V

_ +

12 kΩ 6 kΩ

_ +

*4.54 If υo = −3 V, what is the value of υs in the circuit in Fig. P4.54?

υo 10 kΩ

+ _

3V

+ _

υs _

+

Figure P4.51: Circuit for Problem 4.51.

*4.52

3 kΩ

_ +

7 kΩ

6 kΩ

_ + + _

υo

2V

Find the value of υo in the circuit in Fig. P4.52. Figure P4.54: Circuit for Problem 4.54.

3 kΩ 6 kΩ 8 kΩ

_ +

4 kΩ

_ 9V

+ _

5V

+ 3 kΩ _

Sections 4-9 and 4-10: Instrumentation Amp and D/A Converter

+

8 kΩ

Figure P4.52: Circuit for Problem 4.52.

υo

4.55 The instrumentation-amplifier circuit shown in Fig. 4-23 is used to measure the voltage differential υ = υ2 − υ1 . If the range of variation of υ is from −10 to +10 mV and R1 = R3 = R4 = R5 = 100 k, choose R2 so that the corresponding range of υo is from −5 to +5 V. *4.56 An instrumentation amplifier with R1 = R3 = 10 k, R4 = 1 M, and R5 = 1 k uses a potentiometer for the gaincontrol resistor R2 . If the potentiometer resistance can be varied

“book” — 2015/5/4 — 7:10 — page 243 — #61

PROBLEMS

243

between 10 and 100 , what is the corresponding variation of the circuit gain G = υo /(υ2 − υ1 )? 4.57 Design a five-bit DAC using a circuit configuration similar to that in Fig. 4-25. 4.58 Design a six-bit DAC using a R–2R ladder configuration. Section 4-11: MOSFET

4.61 In Problem 3.73 of Chapter 3, we analyzed a current mirror circuit containing BJTs. Current mirror circuits also can be designed using MOSFETs, as shown in Fig. P4.61. Determine the relationship between I0 and IREF .

IREF

I0

D

4.59 In Example 4-9, we analyzed a common-source amplifier without a load resistance. Consider the amplifier in Fig. P4.59; it is identical to the circuit in Fig. 4-31, except that we have added a load resistor RL . Obtain an expression for υout as a function of υs .

S

D G

S

Figure P4.61: Circuit for Problem 4.61.

VDD RD Section 4-13: Multisim Analysis

Rs

+

D G

+ υs(t) _

S

RL

υout(t)

4.62 Draw a noninverting amplifier (Fig. 4-7) with a gain of 2 in Multisim. Show that the circuit works as expected by connecting a 1 V pulse source and plotting both the input and the output voltages using the Grapher Tool and Transient Analysis. Use the 3-terminal virtual op-amp component.

_

Figure P4.59: MOSFET circuit for Problem 4.59.

*4.60 Determine υout (t) as a function of υs (t) for the circuit in Fig. P4.60. Assume VDD = 2.5 V.

VDD

4.64 In Multisim, draw a summing amplifier that adds the values of four different dc voltage sources, each with an inverting gain of 4. Use the DC Operating Point analysis tool to verify the circuit performance.

1 kΩ 10 Ω + υs(t) _

g1 = 10 A/V

4.63 Draw an inverting amplifier (Fig. 4-11) with a gain of −3.5 in Multisim. Show that the circuit works as expected by connecting a 1 V dc voltage source and solving the circuit using the DC Operating Point analysis. Use the 3-terminal virtual opamp component.

g2 = 100 A/V

+ 1 kΩ

υout(t)

_ Figure P4.60: Two-MOSFET circuit for Problem 4.60.

4.65 In Multisim, draw a noninverting summing amplifier that adds the values of three different dc voltage sources V1 , V2 , and V3 with gains of 1, 2, and 5, respectively. Apply the DC Operating Point Solution tool to demonstrate that the circuit functions as specified. 4.66 Draw the op-amp circuit shown in Fig. P4.66 in Multisim, provide a DC Operating Point Analysis solution that demonstrates its operation, and state what function the circuit performs.

“book” — 2015/5/4 — 7:10 — page 244 — #62

244

CHAPTER 4

50 kΩ

+ υin1

50 kΩ

10 kΩ

_ +

+

50 kΩ

40 kΩ

_

Potpourri Questions

50 kΩ

+ _

υout

+ υin2 _

_

Figure P4.66: Circuit for Problem 4.66.

4.67 Construct the noninverting amplifier circuit shown in Fig. P4.67 in Multisim. Set the value of R to 50 k and then perform a DC Sweep analysis of the input voltage from −5 to +5 V. Plot the Output. Now change the value of R to 80 k and repeat the DC Sweep analysis. Compare the two plots either side by side or by overlapping them using the Overlay Traces button on the Grapher toolbar. (Use the three-terminal virtual op amp for the simulation.)

+ _

Vin

Vout

4.69 Based on the information provided in Table TT9-1 of Technology Brief 9, which types of display technologies are best suited for a large football stadium? A home TV? A cell phone screen? 4.70 What are the limitations of today’s computer memory circuits (ROM and RAM), and what emerging technologies are becoming available to improve them? 4.71 Circuit analysis and design can be performed analytically by applying the techniques covered in this book, or they can be performed by computer simulation. Are these competing or complementary approaches? Explain.

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m4.1

R1

10 kΩ

OPERATIONAL AMPLIFIERS

R

Figure P4.67: Circuit for Problem 4.67.

4.68 Until the 1970s, much research was carried out on analog computers (as distinguished from the digital computers found everywhere today). In fact, analog computers were one of the originally intended users of operational amplifiers. Op amps easily can be incorporated to perform many mathematical operations. Using the basic op-amp circuits shown in this chapter, construct a circuit that expresses the following algebraic equation in voltage:

Ideal Op-Amp Model:

(a) Determine a general expression for υout in terms of the resistor values and is for the circuit of Fig. m4.1 (no Multisim or myDAQ for this part). (b) Find Vout for these specific component values: R1 = 3.3 k, R2 = 4.7 k, R3 = 1.0 k, and Is = 1.84 mA. (c) Replace R2 with a potentiometer. Use myDAQ and the potentiometer to determine Vout for each of the following values of R2 : 2.5 k, 10 k, and 25 k.

R1 R2

_ +

υ = 2x − 3.5y + 0.2z, where υ is the output voltage and x, y, and z are three input voltages. Once you have the circuit designed, build it in Multisim and demonstrate that the circuit behaves appropriately by giving it the following inputs: x = 1.2, y = 0.4, and z = 0.9.

R3

Vout

Is

Figure m4.1 Circuit for Problem m4.1.

“book” — 2015/5/4 — 7:10 — page 245 — #63

PROBLEMS

245

m4.2 Noninverting Amplifier: The circuit in Fig. m4.2 uses a potentiometer whose total resistance is R1 . The movable stylus on terminal 2 creates two variable resistors: βR1 between terminals 1 and 2 and (1−β)R1 between terminals 2 and 3. The movable stylus varies β over the range 0 ≤ β ≤ 1.

υ1 1V

(a) Obtain on expression for G = υo /υs in terms of β. (b) Calculate the amplifier gain for β = 0.0, β = 0.5, and β = 1.0 with component values R1 = 10 k and R2 = 1.5 k.

5

10

t (ms)

−1 V

(c) Let υs be a 100 Hz sinusoidal signal with a 1 V peak value. Plot υo and υs to scale for β = 0.0, β = 0.5, and β = 1.0.

υ2 1V

υs

υ0

+ _ βR1 2

t (ms) 2.5 5 7.5 10

1 R1

(1 − β)R1

Figure m4.3 Input waveforms for Problem m4.3.

3 R2

Figure m4.2 Circuit for Problem m4.2.

m4.3

Summing Amplifier:

(a) Design an op-amp summing circuit that performs the operation υo = −(2.14υ1 + 1.00υ2 + 0.47υ3 ). Use not more than four standard-value resistors with values between 10 k and 100 k. Refer to the resistor parts list in Appendix A of the myDAQ tutorial on the EM . (b) Draw the output waveform υo for the input waveforms υ1 and υ2 shown in Fig. m4.3 and υ3 = 4.7 V. (c) State the minimum and maximum values of υo . m4.4

Signal Processing Circuits:

(a) Design a two-stage signal processor to serve as a “distortion box” for an electric guitar. The first-stage amplifier applies a variable gain magnitude in the range 13.3 to 23.3 while the second-stage amplifier attenuates the signal by 13.3, i.e., the second-stage amplifier has a fixed gain of 1/13.3. Note that when the first-stage amplifier gain is 13.3 the overall distortion box gain is unity. The distortion effect relies on intentionally

driving the first-stage amplifier into saturation (also called “clipping”) when its gain is higher than 13.3. Use a 10 k potentiometer and standard-value resistors in the range 1.0 k and 100 k; see the resistor parts list in Appendix A of the myDAQ tutorial on . You may combine two standard-value resistors in series to achieve the required amplifier gains. (b) Derive a general formula for percent clipping of a unitamplitude sinusoidal test signal; this is the percent of time during one period in which the signal is clipped. The formula includes the peak sinusoidal voltage Vp that would appear at the output of the first-stage amplifier with saturation ignored and the actual maximum value Vs due to saturation. (c) Apply your general formula to calculate percent clipping of a 1 V peak amplitude sinusoidal signal for the potentiometer dial in three positions: fully counterclockwise (no distortion), midscale (moderate distortion), and fully clockwise (maximum distortion). Assume the op-amp outputs saturate at ±13.5 V. (d) Apply a 1 V peak amplitude sinusoidal signal with 100 Hz frequency to the distortion box input and plot its output for the potentiometer dial in the same three positions as above. State the maximum and minimum values of the distortion box output. m4.5 Multiple Op-Amp Stages: Determine Vout in each of the two circuits in Fig. m4.5.

“book” — 2015/5/4 — 7:10 — page 246 — #64

246

CHAPTER 4

OPERATIONAL AMPLIFIERS

R1

_

5.6 kΩ

V1

+

4V

_

V2

+

2V

1

_

10 kΩ

+

+

R3

Vout1

10 kΩ

_ (a) R1

_

5.6 kΩ

V1

+

4V

_

V2

+

2V

1 10 kΩ

_

R2

+

1.5 kΩ

R4

_

R3

1 kΩ

+

+

Vout2

10 kΩ

_

(b) Figure m4.5 Circuits for Problem m4.5.

m4.6 The Importance of Voltage Followers: Suppose you are asked to design a circuit to power a certain gadget and the only source available to you is the 15 V source from your NI myDAQ. Your boss tells you that in order for the gadget to operate properly, its input voltage should be 10.3 V. Moreover, you are told that the input equivalent load resistance of the gadget is exceedingly high (greater than 10 M). To generate the required 10.3 V source, you used the voltage divider shown in Fig. m4.6. (a) Confirm that the voltage divider provides an output voltage of 10.3 V. (b) It turns out that the information given to you about the load resistance is in error; the true load resistance of the gadget is 10 k, not 10 M, and the required input current is 1.03 mA. Reevaluate your circuit in light of the new information. What is the input voltage for the gadget and what is the input current?

+ 15 V _

R1

15 kΩ

R3

33 kΩ

+

V_in

Gadget

Figure m4.6 Circuit for Problem m4.6.

(c) To fix the problem, you decide to use a voltage follower. Design a voltage follower in conjunction with your voltage divider from part (a) to achieve a 1.03 mA current through the 10 k load resistor.

“book” — 2015/5/4 — 7:10 — page 247 — #65

PROBLEMS

247

R6 R3 100 kΩ

+ 1V _

R2

V1

_

1 kΩ

+

R8

5.6 kΩ R1 10 kΩ

1 kΩ

R5 3.3 kΩ R4

+ 1V _

V2

1 kΩ

_ +

V2

Figure m4.7 Circuit for Problem m4.7.

m4.7 Cascaded Op Amps: Find the voltage at each of the three op-amp outputs in the circuit of Fig. m4.7.

3.3 kΩ

R7

V1

+ _

V3

“book” — 2015/5/4 — 7:14 — page 248 — #1

5

CHAPTER

RC and RL First-Order Circuits Contents 5-1 5-2 TB12 5-3 5-4 5-5 TB13 5-6 TB14 5-7 5-8

Overview, 249 Nonperiodic Waveforms, 250 Capacitors, 258 Supercapacitors, 265 Inductors, 269 Response of the RC Circuit, 275 Response of the RL Circuit, 287 Hard Disk Drives (HDD), 293 RC Op-Amp Circuits, 295 Capacitive Sensors, 301 Application Note: Parasitic Capacitance and Computer Processing Speed, 305 Analyzing Circuit Response with Multisim, 310 Summary, 313 Problems, 314

Charge/discharge time

Objectives Learn to: Capacitors (C) and inductors (L) are energy storage devices, in contrast with resistors, which are energy dissipation devices. This chapter examines the behavior of RC and RL circuits, to be followed in Chapter 6 with an examination of RLC circuits.



Use mathematical functions to describe several types of nonperiodic waveforms.



Define the electrical properties of a capacitor, including its i-υ relationship and energy equation.



Combine multiple capacitors when connected in series or in parallel.



Define the electrical properties of an inductor, including its i-υ relationship and energy equation.

Analyze the transient responses of RC and RL circuits.



Combine multiple inductors when connected in series or in parallel.

Design RC op-amp circuits to perform differentiation and integration and related operations.



Apply Multisim to analyze RC and RL circuits.





“book” — 2015/5/4 — 7:14 — page 249 — #2

249

Overview A resistor is characterized by a linear i–υ relationship, namely υ = iR, which does not involve time explicitly. When we apply Kirchhoff’s current and voltage laws to resistive circuits, we end up with one or more simultaneous linear equations. The process of solving a set of linear equations is relatively straightforward and does not involve time explicitly, but if i varies with time, so will υ, in a linearly proportionate manner, and the character of the time variation remains the same for both. Hence, even when a certain voltage or current source in the circuit varies with time, we solve the resistive circuit using static formulas that do not depend on time rather than dynamic formulas that do, because the time variation is merely a scale change. Another important feature of resistive elements is that they consume electrical energy by converting it into heat. Resistive circuits are used to change the relationship between υ and i, divide voltages and currents, and (with the addition of op amps) amplify, add, subtract, and compare voltages. Resistive sensors allow us to convert properties of the physical world—light, heat, sound, moisture, pressure, etc.—to voltage and current values that we can use in our circuits.  Capacitors and inductors represent a contrasting (yet complimentary) class of electrical devices. Not only is time t (or more precisely d/dt) at the heart of how capacitors and inductors function, but they also differ from resistors in that they do not dissipate energy. They can store energy and then release it—but not consume it.  The addition of capacitors and inductors to circuits containing time-varying sources opens the door to dynamic circuits with a wide range of practical applications. Because capacitors and inductors store energy, they can be used to smooth out or average time-varying voltages or currents, select or filter out different frequencies, and delay circuit responses. Capacitive sensors can also be used to measure proximity, touch, pressure, moisture, vibration, and more. Both capacitors and inductors also are found as unintended parasitics in all circuits. The dynamic, time-varying responses of capacitors and inductors provide a new and important set of tools for controlling voltage and current. The dynamic response of a circuit to a certain voltage or current source depends on both the architecture of the circuit and the waveform characterizing the time variation of that source. In general, the

response consists of a transient component and a steady-state component.  The transient response represents the initial reaction immediately after a sudden change, such as closing or opening a switch to connect a source to the circuit. This is also called the early time response. 

Most (but not all) electronic circuits are designed such that the transient response usually dies out or reaches an approximately constant level within a fraction of a second after the introduction of the external excitation. An example of a transient response is when the energy stored in a capacitor is transferred into the flashbulb of a camera. Figure 5-1 shows examples of two typical circuit responses. In part (a), the external excitation is a dc voltage source, and the displayed response represents the current flowing through a certain capacitor in the circuit, starting when the switch is closed. This is much like the camera flash example. The current levels labeled i0 and i∞ denote the values exhibited by the transient response at the onset of the change (closing the switch at t = 0) and a long time afterward (at t = ∞), respectively. They are called the initial and final (or steady state) values of i(t). For this example, the steady state current is i∞ = 0. Our second example displays in Fig. 5-1(b) the response of another circuit to turning on a sinusoidally time-varying source. The combination of the ac source and switch action initially elicit a transient response that quickly transitions into a steady-state response. This steady state ac case belongs to a class of external excitations and circuit responses called periodic waveforms (which repeat periodically). In contrast, a dc waveform is nonperiodic (it does not repeat). As we shall see later, the tools of circuit analysis and design lend themselves to different mathematical approaches when dealing with periodic versus nonperiodic waveforms. We will first examine the behavior of circuits excited by nonperiodic external excitations in this and the following chapter, before we pursue the treatment of periodic ac circuits starting in Chapter 7. Section 5-1 introduces some of the nonperiodic waveforms commonly used in electric circuits, followed in Sections 5-2 and 5-3 with presentations of the circuit properties of capacitors and inductors, respectively. Our treatment of the circuit response to nonperiodic excitations is divided into two segments. The first, covered in Sections 5-4 through 5-6 of this chapter, deals with first-order circuits, so named because their Kirchhoff voltage and current equations are characterized by first-order differential equations.

“book” — 2015/5/4 — 7:14 — page 250 — #3

250

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

i(t) i0 (initial value)

υ(t)

+ _

i (final value) t 8

R

0 t=0

(a) dc transient response Circuit

i(t)

Transient response Steady−state response

t

0

(b) Combined response to ac excitation Figure 5-1: Circuit response to (a) dc source υ(t) = V0 and (b) ac source υ(t) = V0 cos ωt.

 First-order circuits include RC circuits—composed of sources, resistors, and a single capacitor (or multiple capacitors that can be combined into a single equivalent capacitor)—and RL circuits, but not circuits containing capacitors and inductors simultaneously. 

RLC circuits, which give rise to second-order differential equations, are the subject of Chapter 6.

Concept Question 5-2: Why do we study the circuit

response to dc and ac sources separately? (See

5-1

)

Nonperiodic Waveforms

Among the multitudes of possible nonperiodic waveforms, the step, ramp, pulse, and exponential waveforms are encountered most frequently in electrical circuits. In this section, we review the geometrical properties and corresponding mathematical expressions associated with each of these four waveforms, as well as introduce some of the connections between them.

Concept Question 5-1: What is the difference between

5-1.1

the transient and steady-state components of the circuit response? (See )

Step-Function Waveform

The waveform υ(t) shown in Fig. 5-2(a) is an (ideal) step function: it is equal to zero for t < 0, at t = 0 it makes a

“book” — 2015/5/4 — 7:14 — page 251 — #4

5-1

NONPERIODIC WAVEFORMS

251 where u(t) is known as the unit step function and is defined as

Step Functions



υ(t) u(t) =

V0 u(t)

V0

0 (a) Ideal step function

for t < 0, for t > 0.

(5.2)

In reality, it is not possible to turn on a switch with an (ideal) step function, because that would require changing the value of υ(t) from 0 to V0 in zero time. A more realistic shape of the step function is illustrated in Fig. 5-2(b); the discontinuous jump is replaced with a ramp waveform with rise time t, providing a smooth voltage turn-on. If υ(t) transitions between its two levels at a time other than zero, such as at t = T , it is written as

t

υ(t) V0

∆t (b) Realistic step function

0 1



t υ(t) = V0 u(t − T ) =

0 V0

for t < T , for t > T .

(5.3)

υ(t) V0 u(t − 3)

V0

 u(t −T ) is called the time-shifted step function, which is defined to be zero when its argument (t −T ) is less than zero and 1 when its argument is greater than zero. Thus, u(t − T ) = 1 for t > T . 

T>0 t (s) 0 3 (c) Time-shifted step u(t − T) with T = 3s υ(t) V0 u(3 − t)

V0

t (s) 0 3 (d) Time-shifted step u(t − T) with T = 3s Figure 5-2: Step functions: (a) ideal step function, (b) realistic step function with transition duration t, (c) time-shifted step function V0 u(t − 3), (d) time-shifted step function V0 u(3 − t).

discontinuous jump to V0 , and from there on forward it remains at V0 . The process represents an ideal switch that turns on a dc voltage at t = 0. Mathematically, it can be described as υ(t) = V0 u(t),

(5.1)

By the same definition, u(T − t) is zero when T − t < 0 (which is true when t > T ), and 1 when T − t > 0 (which is true when t < T ). Figure 5-2(c) and (d) display step-function waveforms for V0 u(t − 3) and V0 u(3 − t), respectively. We often use combinations of step functions to represent voltage sources turning on and off. An example of a step function is when a switch is closed so as to connect a voltage source to a circuit, as shown in Fig. 5-3(a). When writing KCL and KVL equations for circuits that include switches, the switching action (closing or opening) can be represented mathematically by step functions. In Fig. 5-3(a), closing the switch at t = 3 s is represented by u(t − 3), whereas disconnecting the source by opening the switch in Fig. 5-3(b) is represented by u(3 − t). If the time associated with closing the switch is very short in comparison with the time scale of interest, then it may be acceptable to approximate the switch closing by an ideal step function. On the other hand, if we are interested in analyzing the circuit response at a sampling rate whose interval is shorter than or comparable with the transition interval associated with closing the switch, then it may be necessary to use a more realistic, continuous, step function to represent the switch action.

“book” — 2015/5/4 — 7:14 — page 252 — #5

252

CHAPTER 5

t=3s a V0

+ _

RC AND RL FIRST-ORDER CIRCUITS

Ramp Functions a

t=3s Circuit

V0

+ _

υ(t)

Circuit

3V 2V

r(t + 1)

1V

b −3 −2 −1 0

1

2

3

4

t (s)

(a)

a

a υ(t)

+ _

+ _

Circuit

Circuit

3V

V0 u(3 − t)

V0 u(t − 3) b

2V

(a) Switch closes at t = 3 s

r(t − 2)

1V

b

−3 −2 −1 0

(b) Switch opens at t = 3 s

1

2

3

4

t (s)

(b)

Figure 5-3: Connecting/disconnecting a voltage source to/from

υ(t) Slope = 3 V/s

a circuit via a switch can be represented mathematically by a step function.

6V 3r(t − 1)

3V

5-1.2

Ramp-Function Waveform

−3 −2 −1 0

A waveform that varies linearly with time, starting at a specific time t = T , is called a time-shifted ramp function and is denoted by r(t − T ). If T = 0, it simply is called a ramp function and is denoted by r(t). Formally, r(t − T ) is defined as 

0 r(t − T ) = (t − T )

for t ≤ T , for t ≥ T .

(5.4)

1

2

3

4

t (s)

(c) υ(t) 2V t (s) −3 −2 −1 0 1 2 3 4 Slope = −2 V/s −2 V −2r(t + 1) −4 V (d)

Plots of υ(t) = r(t − T ) are displayed in Fig. 5-4(a) and (b) for T = −1 s and T = 2 s, respectively. A voltage υ(t) that ramps up at 3 V per second, starting at t = 1 s, is shown graphically in Fig. 5-4(c). Mathematically, υ(t) can be expressed as

Figure 5-4: Time-shifted ramp functions.

A unit ramp function is related to the unit step function by υ(t) = 3r(t − 1)

V.

(5.5)

If the coefficient of r(t − T ) is negative, υ(t) would exhibit a negative slope, as illustrated by Fig. 5-4(d) for υ(t) = −2r(t + 1).

t r(t) =

u(t) dt = t u(t), −∞

(5.6)

“book” — 2015/5/4 — 7:14 — page 253 — #6

5-1

NONPERIODIC WAVEFORMS

253

and for the case where the ramping action starts at t = T , t r(t − T ) =

u(t − T ) dt = (t − T ) u(t − T ).

(5.7)

−∞

and a second ramp function that starts at T = +2 ms but its slope is −3 V/s. Thus, υ(t) = υ1 (t) + υ2 (t) = 3r(t + 2 ms) − 3r(t − 2 ms)

V.

In view of Eq. (5.7), υ(t) also can be expressed in terms of time-shifted step functions as υ(t) = 3(t + 2 ms) u(t + 2 ms)

Example 5-1: Realistic Step Waveform

− 3(t − 2 ms) u(t − 2 ms) Generate an expression to describe the waveform shown in Fig. 5-5(a). Note that the time scale is in ms. Solution: The voltage υ(t) can be synthesized as the sum of two time-shifted ramp functions (Fig. 5-5(b)): one with a positive slope of 3 V/s and a ramp start-up time T = −2 ms

υ (V) 12 9 6

5-1.3

V.

Pulse waveform

The diagram in Fig. 5-6(a) depicts a SPDT switch that moves from position 1 to position 2 at t = 1 s, connects a dc voltage source to an electric circuit, and then returns to position 1 at t = 5 s. From the standpoint of the circuit, the switch actions constitute the introduction of a rectangular pulse of voltage V0 , as illustrated in Fig. 5-6(b). A pulse also may be triangular or Gaussian in shape or may assume other forms, but in all cases, it usually is assumed that a pulse rises from some specified base level up to a peak value, remains constant for a while, and then declines back to its original base level.

3 0 1

−5 −4 −3 −2 −1

2

3

t (ms)

4

Moves from 1 to 2 @ t = 1 s Returns to 1 @ t = 5 s

(a) Original function Composite waveform

υ (V)

V0

+ _

SPDT 2

1

+ υ(t)

_

Circuit

12 9

υ1(t) = 3r(t + 2 ms)

(a) Circuit with input switch

6 3 −5 −4 −3 −2 −1 0 1 −3

2

3

4

t (ms)

−6 −9 −12

+

V0

υ(t) t=1s

t=5s

_

Circuit

3s υ2(t) = −3r(t − 2 ms)

(b) As sum of two time-shifted ramp functions Figure 5-5: Step waveform of Example 5-1.

(b) Equivalent  inputpulse t −3 rect 4 Figure 5-6: Connecting a switch to a dc source at t = 1 s and

then returning it to ground at t = 5 s constitutes a voltage pulse centered at T = 3 s and of duration τ = 4 s.

“book” — 2015/5/4 — 7:14 — page 254 — #7

254

CHAPTER 5

A rectangular pulse can be constructed out of two time-shifted step functions: one that causes the rise in level and another (delayed in time) that cancels the first one. The details are given in Example 5-2.

Rectangular Pulses υ(t) rect τ

1

t−T τ

( )

Example 5-2: Pulses

0

t (s)

T

Construct expressions for (a) the rectangular pulse shown in Fig. 5-8(a) and (b) the trapezoidal pulse shown in Fig. 5-8(b) in terms of step and ramp functions.

(a)

Solution: (a) From Fig. 5-8(a), it is evident that the amplitude of the rectangular pulse is 4 V and its duration is 2 s, extending from T1 = 2 s to T2 = 4 s. Hence, with its center at 3 s and its duration equal to 2 s,

υ(t) V0 rect

RC AND RL FIRST-ORDER CIRCUITS

t+2 2

( )

V0

2 −3 −2 −1 T = −2



t (s)

0

υa (t) = 4 rect

υ(t)

2 −8

 V.

(5.9)

The sequential addition of two time-shifted step functions, υ1 (t) at t = 2 s and υ2 (t) at t = 4 s, as demonstrated graphically in Fig. 5-8(c), accomplishes the task of synthesizing the rectangle function in terms of two step functions. Specifically,

(b)

0

t −3 2

3

T=3 4

2 −8 rect

υa (t) = υ1 (t) + υ2 (t) = 4[u(t − 2) − u(t − 4)]

t (s)

V. (5.10)

(b) The trapezoidal pulse consists of three segments, a ramp with a positive slope that starts at t = 0 and ends at t = 1 s, followed by a plateau that extends to t = 3 s, and finally, a ramp with a negative slope that ends at 4 s. Building on the experience gained from Example 5-1, we can synthesize the trapezoidal pulse in terms of four ramp functions. The process, which is illustrated graphically in Fig. 5-8(d), leads to

t−3 2

( )

(c) Figure 5-7: Rectangular pulses.

υb (t) = υ1 (t) + υ2 (t) + υ3 (t) + υ4 (t) A rectangular pulse can be described in terms of the unit rectangular function rect[(t − T )/τ ], which is characterized by two parameters: location of the center of the pulse T and the duration of the pulse τ , as shown in Fig. 5-7. Its mathematical definition is given by

V. (5.11)

Equivalently, using the relationship between the ramp and step functions given by Eq. (5.7), υb (t) can be expressed as υb (t) = 5[t u(t) − (t − 1) u(t − 1)



 t −T rect τ ⎧ ⎪ ⎨0 for t < (T − τ/2), = 1 for (T − τ/2) ≤ t ≤ (T + τ/2), ⎪ ⎩ 0 for t > (T + τ/2).

= 5[r(t) − r(t − 1) − r(t − 3) + r(t − 4)]

− (t − 3) u(t − 3) + (t − 4) u(t − 4)]

(5.8)

V. (5.12)

There are often multiple ways for representing waveforms of these types, all of which should lead to the same result in the end.

“book” — 2015/5/4 — 7:14 — page 255 — #8

5-1

NONPERIODIC WAVEFORMS

255

Waveform Synthesis υa(t)

υb(t)

t−3 4 rect 2

( )

5V

4V 0 1

2

3

4

5

t (s)

0 1

−2 −1

υb(t)

υa(t) 4u(t − 2)

−4 V

3 4

4

5

t

5

υ1(t) υ4(t)

5V

4V 2

3

(b) Trapezoidal pulse

(a) Rectangular pulse

0 1

2

t (s)

0 1

−2 −1

−4u(t − 4)

2

3

4

t

υ3(t)

υ2(t)

(c) υa(t) = 4u(t − 2) − 4u(t − 4)

5

(d) υb(t) = υ1(t) + υ2(t) + υ3(t) + υ4(t)

Figure 5-8: Rectangular and trapezoidal pulses of Example 5-2.

υ

Concept Question 5-3: What determines the slope of a

ramp waveform? (See

)

10

0

Concept Question 5-4: How are the ramp and rectangle

functions related to the step function? (See

)

2

t (s)

4

−10 (a) υ

Concept Question 5-5: A unit step function u(t) is

equivalent to closing an SPST switch at t = 0. What is u(−t) equivalent to? (See )

5

0 Exercise 5-1: Express the waveforms shown in Fig. E5.1

in terms of unit step functions. Answer:

(a) υ(t) = 10 u(t) − 20 u(t − 2) + 10 u(t − 4), (b) υ(t) = 2.5 r(t) − 10 u(t − 2) − 2.5 r(t − 4). (See )

2

4

−5 (b) Figure E5.1

t (s)

“book” — 2015/5/4 — 7:14 — page 256 — #9

256

CHAPTER 5

Exercise 5-2: How is u(t) related to u(−t)? Answer: They are mirror images of one another (with

respect to the y-axis). (See

C

Assume that it started out at position 2, was moved to position 1 at t = 1 s, and then moved back to position 2 at t = 5 s. This is the reverse of the sequence shown in Fig. 5-6(a). Express υ(t) in terms of (a) units step functions and (b) the rectangle function. Answer: (a) υ(t) = V0 [u(1 − t) + u(t − 5)], (b) υ(t) = V0 1 − rect t−3 . (See ) 4

Exponential waveform

The exponential function is a particularly useful tool for characterizing fast-rising and fast-decaying waveforms, which, as we will see in later sections, are related to the transient responses of RC and RL circuits. The (positive) exponential function given by υp (t) = et/τ

(5.13)

is shown graphically in Fig. 5-9 for a positive value of the time constant τ . The figure also includes a plot of the negative exponential function, where υn (t) = e

time, it means that after τ seconds its amplitude decreases to 1/e or 37 percent of its initial value. Symmetrically, υp = e−1 = 0.37 when t = −τ .

)

Exercise 5-3: Consider the SPDT switch in Fig. 5-6(a).

5-1.4

RC AND RL FIRST-ORDER CIRCUITS

−t/τ

.

(5.14)

When t = τ , υn = e−1 = 0.37. Thus, if a certain quantity (such as a voltage or current) is said to decay exponentially with

Positive exponential

 An exponential function with a short time constant rises or decays faster than an exponential function with a longer time constant, as illustrated by the plots in Fig. 5-10(a). 

Replacing t in the exponential with (t −T ) shifts the exponential curve to the right if T has a positive value and to the left if T is negative (Fig. 5-10(b)). In Fig. 5-10(c), the range of the exponential function has been limited to t > 0 by multiplying e−t/τ by u(t), and in Fig. 5-10(d) the function υ(t) = V0 (1 − e−t/τ ) u(t) is used to describe a waveform that builds up as a function of time towards a saturation value V0 . Table 5-1 provides a summary of common waveform shapes and their equivalent expressions. Concept Question 5-6: If the time constant of a

negative exponential function is doubled in value, will the corresponding waveform decay faster or slower? (See ) Concept Question 5-7: What is the approximate shape

of the waveform described by the function (1 − e−|t|)? (See )

Exercise 5-4: The radioactive decay equation for a certain

material is given by n(t) = n0 e−t/τ , where n0 is the initial count at t = 0. If τ = 2 × 108 s, how long is its half-life? [Half-life t1/2 is the time it takes a material to decay to 50 percent of its initial value.] Answer: t1/2 = 1.386 × 108 s = 4 years, 144 days,

υp = et/τ

12 hours, 10 minutes, 36 s. (See

1

)

υn = e−t/τ 0.37 −3

−2

−1

0

1

Negative exponential t/τ 2 3

Figure 5-9: By t = τ , the exponential function e−t/τ has decayed to 37 percent of its original value at t = 0.

Exercise 5-5: If the current i(t) through a resistor R decays exponentially with a time constant τ , what is the value of the power dissipated in the resistor at t = τ , compared with its value at t = 0? Answer: p(t) = i 2 R = I02 R(e−t/τ )2 = I02 Re−2t/τ ,

p(τ )/p(0) = e−2 = 0.135 or 13.5 percent. (See

C3

)

“book” — 2015/5/4 — 7:14 — page 257 — #10

5-1

NONPERIODIC WAVEFORMS

257

Exponential Functions υ(t)

e−t/2 e−t

et

υ(t)

et/2

e(t + 1) et (t − 1) e

Longer time constant, slower decay Shorter time constant,

1

1

faster decay t

0

t

0 −1 1 (b) Role of time shift T

(a) Role of time constant τ υ(t)

υ(t)

V0

V0 0.63V0

V0e−t/τ u(t)

0.37V0

t

τ

0

V0[1 − e−t/τ] u(t)

(c)

t

τ

0 (d)

Figure 5-10: Properties of the exponential function.

Table 5-1: Common nonperiodic waveforms. waveform Step

Ramp

Expression  0 u(t − T ) = 1

for t < T for t > T

General Shape

u(t − T) 1 0 T r(t − T)

r(t − T ) = (t − T ) u(t − T )

0 T 

Rectangle

rect



t −T = u(t − T1 ) − u(t − T2 ) τ τ τ T1 = T − ; T2 = T + 2 2 rect

1 0 T1

Slope = 1 t

t− T τ t

T2

exp[−(t − T)/τ] u(t − T)

1 Exponential

t

exp[−(t − T )/τ ] u(t − T )

0

T

t

“book” — 2015/5/4 — 7:14 — page 258 — #11

258

CHAPTER 5

+ + + υ

+ _

E _

+ +

+

_ _

+ +

+

Table 5-2: Relative electrical permittivity of common

Area A

+q

insulators: εr = ε/ε0 and ε0 = 8.854 × 10−12 F/m.

+ +

Air (at sea level) Teflon Polystyrene Paper Glass Quartz Bakelite Mica Porcelain

E _ Dielectric ε −q

Figure 5-11: Parallel-plate capacitor with plates of area A, separated by a distance d, and filled with an insulating dielectric material of permittivity ε.

5-2 Capacitors  When separated by an insulating medium, any two conducting bodies (regardless of their shapes and sizes) form a capacitor. A capacitor can store electric charge.  The parallel-plate capacitor shown in Fig. 5-11 represents a simple configuration in which two identical conducting plates (each of area A) are separated by a distance d containing an insulating (dielectric) material of electrical permittivity ε. The permittivity of a material is usually referenced to that of free space, namely ε0 = 8.85 × 10−12 farads/m (F/m). Hence, the relative permittivity of a material is defined as ε . ε0

(5.15)

When a dielectric material is subjected to an electric field, its atoms become partially polarized; i.e., the atom is rearranged into positive and negative domains.. The electric field E induced in the space between the conducting plates is the result of the voltage υ applied across the plates. The electrical susceptibility χe of a material is a measure of how susceptible that material is to electrical polarization. The permittivity ε and susceptibility χe are related by ε = ε0 (1 + χe ).

Relative Permittivity εr

Material

d

+

_ _ _ _ _ _ _ _

εr =

RC AND RL FIRST-ORDER CIRCUITS

(5.16)

1.0006 2.1 2.6 2–4 4.5–10 3.8–5 5 5.4–6 5.7

Free space contains no atoms; hence, its χe = 0 and εr = 1. For air at sea level, εr = 1.0006 ≈ 1.0. Table 5-2 provides typical values of εr for common types of insulators. Returning to the parallel-plate capacitor, if a voltage source is connected across the two plates, as shown in Fig. 5-11, charge of equal and opposite polarity is transferred to the conducting surfaces. The plate connected to the (+) terminal of the voltage source will accumulate charge +q, and charge −q will accumulate on the other plate. The charges induce a nearly uniform electric field E in the dielectric medium, given by E=

q , εA

(5.18)

with the direction of E being from the plate with +q to the plate with −q. Moreover, E, whose unit is V/m, is related to the voltage υ through E=

υ d

(V/m)

(parallel-plate capacitor).

(5.19)

 For any capacitor, its capacitance C, measured in farads (F), is defined as the amount of charge q that its positive-polarity plate holds, normalized to the applied voltage responsible for that charge accumulation.  Thus,

In view of Eq. (5.15), the relative permittivity εr is given by εr =

ε = 1 + χe . ε0

(5.17)

C=

q υ

(F)

(any capacitor).

(5.20)

“book” — 2015/5/4 — 7:14 — page 259 — #12

5-2

CAPACITORS

259

For the parallel-plate capacitor, combining Eqs. (5.18) and (5.19) leads to q = εAυ/d. Upon inserting this expression for q in Eq. (5.20), we have

Conductors C=

εA d

(parallel-plate capacitor).

Even though the expression given by Eq. (5.21) is specific to the parallel-plate capacitor, the general tenor of the expression holds true for other geometrical configurations as well. In general, the capacitance C of any two-conductor system increases with the area of the conducting surfaces, decreases with the separation between them, and is directly proportional to ε of the insulating material. For example, the capacitance of a coaxial capacitor consisting of two concentric conducting cylinders of radii a and b (Fig. 5-12(a)) and separated by a dielectric material of permittivity ε is given by

C=

2πε ln(b/a)

(coaxial capacitor),

l

(5.21)

Dielectric ε

2a 2b (a) Coaxial capacitor

Metal foil Mica insulator

(5.22)

(b) Mica capacitor where is the length of the capacitor and ln(b/a) is the natural logarithm of (b/a). The spacing between the cylinders is (b−a); reducing this spacing, while holding b constant, requires reducing the ratio (b/a), which reduces the value of ln(b/a), thereby increasing the magnitude of C. The mica capacitor shown in Fig. 5-12(b) consists of a stack of conducting plates, interleaved by sheets of mica (dielectric). The plastic-foil capacitor in Fig. 5-12(c) is constructed by rolling flexible conducting foils (separated by a plastic layer) into a spindle-like configuration. Small capacitors used in microcircuits typically have capacitances in the picofarad (10−12 F) to microfarad (10−6 F) range. Large capacitors used in power-transmission substations may have capacitors in the range of millifarads (10−3 F). Using thin-film polymers for the dielectric insulator and carbon nanotubes for the electrodes (terminals), a new type of capacitor (sometimes called a supercapacitor or nanocapacitor) was developed in the 1990s with the express goal of significantly increasing the amount of charge that the conductors can hold (at a specified voltage level). Such capacitors have capacitance values that are several orders of magnitude greater than conventional capacitors of comparable size. The new fabrication techniques have not only expanded the versatility of capacitors in electronic circuits, but they have also introduced the use of supercapacitors as energy-storage devices in many electronic applications (see Technology Brief 12: Supercapacitors).

Lead to inner foil sheet

Inner metal foil Outer metal foil

Lead to outer foil sheet

Plastic insulator

(c) Plastic foil capacitor Figure 5-12: Various types of capacitors.

5-2.1

Electrical Properties of Capacitors

According to Eq. (5.20), q = Cυ. Application of the standard definition for current (Eq. (1.3) provides the expression for the current i through a capacitor as i=

dq dυ =C , dt dt

(5.23)

where the direction of i and the polarity of υ are defined in accordance with the passive sign convention (Fig. 5-13).

“book” — 2015/5/4 — 7:14 — page 260 — #13

260

CHAPTER 5

i C

+ _υ

i=C

dυ dt

Figure 5-13: Passive sign convention for capacitor: if current i

t In view of dq = i dt, we recognize that the integral t0 i dt represents the amount of charge accumulation on the capacitor at time t. If we are dealing with a capacitor that had no charge on it until a switch was closed or a signal was injected into the circuit and if we conveniently set our time reference such that the signal injection commenced at t0 = 0, then Eq. (5.25) simplifies to

is entering the (+) voltage terminal across the capacitor, then power is getting transferred into the capacitor. Conversely, if i is leaving the (+) terminal, then power is getting released from the capacitor.

The i–υ relationship expressed by Eq. (5.23) conveys a very important condition, namely:  The voltage across a capacitor cannot change instantaneously, but the current can.  This assertion is supported by the observation that if υ were to change values in zero time, dυ/dt would be infinite, as a result of which the current i would be also infinite. Since i cannot be infinite, υ cannot change instantaneously. Another attribute of Eq. (5.23) relates to the behavior of a capacitor under dc conditions (constant voltage across it). Since dυ/dt = 0 for a dc voltage, it follows that i = 0. Such a behavior is characteristic of an open circuit, through which no current flows even when a non-zero voltage exists across it. Thus:  Under dc conditions, a capacitor behaves like an open circuit.  To express υ(t) in terms of i(t), we replace t with a dummy variable t  and integrate both sides of Eq. (5.23) from t  = t0 to t  = t, t   t dυ 1  dt = i dt  , (5.24) dt  C t0

t0

where t0 is the initial reference point in time at which the initial condition υ(t0 ) is known. Since the integral of the derivative of a function is the function itself, integrating the left-hand side and rearranging terms leads to

υ(t) = υ(t0 ) +

1 C

t t0

1 υ(t) = C

(5.25)

t

i dt 

(5.26)

0

(capacitor uncharged before t = 0).

Charging up a capacitor creates an electric field in the dielectric medium between the capacitor’s conductors. The electric field becomes the mechanism for storage of electrical energy in that medium. The stored energy can be released by discharging the capacitor. Thus, a capacitor can store energy and release previously stored energy but cannot dissipate energy. The instantaneous power p(t) transferring into or out of a capacitor is given by dυ dt

p(t) = υi = Cυ

(W),

(5.27)

where i is defined as entering the capacitor at its positive voltage terminal (Fig. 5-13).  If the magnitude of p(t) is positive, then by the passive sign convention, the capacitor is receiving power (charging up), and if p(t) is negative, it is delivering power (discharging).  Energy is the integral of the product of power and time. Hence, the amount of energy stored in the capacitor at any time t is equal to the time integral of p(t) from −∞ (at which time the capacitor was uncharged) to t and is given by t

t 



υ

p dt = C

w(t) = −∞

i dt  .

RC AND RL FIRST-ORDER CIRCUITS

−∞

t  =C −∞

dυ dt 

d dt 





dt 

1 2 υ 2



dt  ,

(5.28)

“book” — 2015/5/4 — 7:14 — page 261 — #14

5-2

CAPACITORS

261

which yields

υ (V) w(t) =

1 C υ 2 (t) 2

(J).

We note that since the capacitor had no charge at t = −∞, then its voltage also was zero at t = −∞. Equation (5.29) states that:

Voltage

10 5

(5.29)

0 1

t (s) 2

3

4

5

6

7

5

6

7

(a) i (μA)

 The electrical energy stored in a capacitor at a given instant in time depends on the voltage across the capacitor at that instant, without regard to prior history. 

6 3

This stored energy is akin to potential energy in a physical system.

−3

Current 0 1

t (s) 2

3

4

(b) Example 5-3: Capacitor Response to Voltage Waveform

p (μW)

The voltage waveform shown in Fig. 5-14(a) was applied across a 0.6 μF capacitor. Determine the corresponding waveforms for (a) the current i(t), (b) the power p(t), and (c) the energy stored in the capacitor w(t). Solution: (a) We start by establishing a suitable expression for the waveform of υ(t), shown in Fig. 5-14(a), in terms of ramp functions. Noting that the ramp starts at t = 0 and has a slope of 10/2 = 5 V/s, υ(t) can be written as υ(t) = 5r(t) − 5r(t − 2) − 5r(t − 4) + 5r(t − 5)

−15 −30

0 1

2

3

4

5

(c)

t (s) 6 7 Power transfer out of capacitor (discharging)

w (μJ)

V.

30 22.5 15 7.5

Recalling that according to Eq. (5.7), r(t − T ) = (t − T ) u(t − T ), the expression for υ(t) corresponds to ⎧ ⎪ 0 for t ≤ 0, ⎪ ⎪ ⎪ ⎪ ⎪ 5t V for 0 ≤ t ≤ 2 s, ⎨ υ(t) = 10 V for 2 s ≤ t ≤ 4 s, ⎪ ⎪ ⎪ (−5t + 30) V for 4 s ≤ t ≤ 5 s, ⎪ ⎪ ⎪ ⎩5 V for t ≥ 5 s.

Power transfer into capacitor (charging)

30 15

Energy

0

1

2

3

4

5

6

7

t (s)

(d) (5.30)

Application of Eq. (5.23), while recalling that the derivative is the same as the slope of a line or curve, gives: ⎧ ⎪ 0 for t ≤ 0, ⎪ ⎪ ⎪ ⎪ ⎪ for 0 ≤ t ≤ 2 s, ⎨3 μA dυ i(t) = C = 0 (5.31) for 2 s ≤ t ≤ 4 s, ⎪ dt ⎪ ⎪ −3 μA for 4 s ≤ t ≤ 5 s, ⎪ ⎪ ⎪ ⎩0 for t ≥ 5 s.

Figure 5-14: Example 5-3 waveforms for i, υ, p, and w.

A plot of the current waveform is displayed in Fig. 5-14(b). We note that i(t) > 0 when υ(t) has a positive slope, and i(t) < 0 when υ(t) has a negative slope. (b) The power p(t), which is equal to the product of Eqs. (5.30) and (5.31), is shown in Fig. 5-14(c). (c) We can calculate the stored energy w(t) either by integrating p(t)—which is graphically equivalent to computing the area under the curve—or by applying Eq. (5.29). In either case, we end up with the plot displayed in Fig. 5-14(d).

“book” — 2015/5/4 — 7:14 — page 262 — #15

262

CHAPTER 5

We note that after t = 5 s, the current is zero, the voltage is constant, the power getting transferred into the capacitor is zero (because i = 0), and the stored energy remains unchanged at 7.5 μJ. Let us examine the energy transfer process from the standpoint of the current and voltage. Between t = 0 and 2 s, a constant positive current flows to the capacitor, causing the deposition of positive charge on one side of the capacitor and a net increase of negative charge by the same amount on the other side of the capacitor. The increase in charge leads to a linear increase in voltage. By Eq. (5.29), increasing the voltage leads to a quadratic increase in stored energy, as shown in Fig. 5-14 during the time span between 0 and 2 s. Between 2 and 4 s, i = 0 and υ is a constant. Hence, the stored energy remains unchanged. Then, between 4 and 5 s, the current reverses direction, which entails repatriating some of the positive charges back to their original location. Consequently, υ decreases and so does the stored energy, until t = 5 s. Beyond that time, the remaining charge stays in place, the voltage remains constant at 5 V, and the corresponding 7.5 μJ of energy stored in the capacitor remains in that state until some future action. Example 5-4: RC Circuit under dc Conditions

RC AND RL FIRST-ORDER CIRCUITS

30 kΩ 20 kΩ 20 V

+ _

C1

+ _ V1

C2

+ _ V2 50 kΩ 40 kΩ

(a) Original circuit

V 30 kΩ 20 kΩ 20 V

+ _

C1

+ _V1

C2

+ _V2 50 kΩ 40 kΩ

(b) Equivalent circuit Figure 5-15: Under dc conditions, capacitors behave like open circuits.

Determine voltages υ1 and υ2 across capacitors C1 and C2 in the circuit of Fig. 5-15(a). Assume that the circuit has been in its present (charged) condition for a long time. Solution: “Long time” implies steady state. Under steadystate dc conditions, no current flows through a capacitor. Replacing capacitors C1 and C2 with open circuits, as in Fig. 5-15(b), allows us to apply KCL at node V as V V − 20 + = 0, 3 20 × 10 (30 + 50) × 103

cannot change instantaneously. Can the current change instantaneously, and why? (See )

Concept Question 5-10: For the capacitor, can p(t) be negative? Can w(t) be negative? Explain. (See )

which gives V = 16 V. Hence, V1 = V = 16 V. Through voltage division, V2 across the 50 k resistor is given by V × 50k 16 × 50 = = 10 V. V2 = (30 + 50)k 80 Concept Question 5-8: Explain why a capacitor behaves

like an open circuit under dc conditions. (See

Concept Question 5-9: The voltage across a capacitor

)

Exercise 5-6: It is desired to build a parallel-plate capacitor capable of storing 1 mJ of energy when the voltage across it is 1 V. If the capacitor plates are 2 cm × 2 cm each and its insulating material is Teflon, what should the separation d be? Is such a capacitor practical? Answer: d = 3.72 × 10−12 m. No, it is not practical to

build a capacitor with such a small d, because it is about two orders of magnitude smaller than the typical spacing between two adjacent atoms in a solid material. C3) (See

“book” — 2015/5/4 — 7:14 — page 263 — #16

5-2

CAPACITORS

263

Exercise 5-7: Instead of specifying A and calculating

the spacing d needed to meet the 1 mJ requirement in Exercise 5-6, suppose we specify d as 1 μm and then calculate A. How large would A have to be?

Answer: A = 10.4 m × 10.4 m, equally impractical!

(See

C3

)

Combining In-Series Capacitors is

1

υ1 + _

υ2 + _

υ3 + _

C1

C2

C3

+ υs + _ 2

Exercise 5-8: Determine the current i in the circuit of

Fig. E5.8, under dc conditions.

is

1 μF 5 kΩ

+ υs + _

15 kΩ 2 μF

1 Equivalent circuit

i

 Ceq =

 1 1 1 −1 + + C1 C2 C3

2

40 kΩ

1.5 A

20 kΩ

Figure 5-16: Capacitors in series.

Figure E5.8 Answer: i = 1 A. (See

5-2.2

+ _

C

)

Series and Parallel Combinations of Capacitors

In Chapter 2, we established that multiple resistors connected in series are equivalent to a single resistor whose resistance is equal to the algebraic sum of the resistances of the individual resistors. This equivalence relationship does not hold true for capacitors. In fact, we will shortly determine that:  The equivalence relationship for capacitors connected in series is similar in form to the relationship for resistors connected in parallel, and vice versa. 

We wish to relate Ceq of the equivalent circuit to C1 , C2 , and C3 , subject to the requirement that the actual circuit and its equivalent exhibit identical i–υ characteristics at terminals (1, 2). For the equivalent circuit,   dυ1 dυs dυ2 dυ3 = Ceq + + is = Ceq dt dt dt dt   is is is , (5.34) = Ceq + + C1 C2 C3 which leads to 1 1 1 1 = + + . Ceq C1 C2 C3

(5.35)

Generalizing to the case of N capacitors in series, N

 1 1 1 1 1 = = + + ··· + Ceq Ci C1 C2 CN

Capacitors in series Consider the three capacitors shown in Fig. 5-16. They share the same current is , and are therefore in series. Current is related to their individual voltages by dυ1 dυ2 dυ3 = C2 = C3 . is = C1 dt dt dt

(5.32)

Also, υs = υ1 + υ2 + υ3 .

(5.33)

(5.36)

i=1

(capacitors in series). Additionally, if at reference time t0 the capacitors had initial voltages υ1 (t0 ) to υN (t0 ), the initial voltage of the equivalent capacitor is υeq (t0 ) =

N  i=1

υi (t0 ).

(5.37)

“book” — 2015/5/4 — 7:14 — page 264 — #17

264

CHAPTER 5

Example 5-5: Equivalent Circuit

Combining In-Parallel Capacitors is

υs

1

+ _

i1

i2

i3

C1

C2

C3

Reduce the circuit of Fig. 5-18(a) into the simplest equivalent configuration. Solution: Resistors are combined independently of capacitors. For the resistors, we first combine R2 and R3 in parallel, and then add the result to R1 in series, noting that interchanging the locations of two elements connected in series is perfectly permissible, as such an action has no influence on either the current flowing through them or the voltages across them. A similar procedure can be followed for the capacitors, but we have to keep in mind that the equivalence relationships for resistors and capacitors are the reciprocal of one another:

2

is υs

+ _

1 Equivalent circuit

RC AND RL FIRST-ORDER CIRCUITS

Ceq = C1 + C2 + C3

R2  R3 =

R2 R 3 3k × 6k = 2 k . = R2 + R 3 3k + 6k

Req = R1 + 2 k = 8 k + 2 k = 10 k ,

2

C2  C3 = C2 + C3 = 1 μF + 5 μF = 6 μF,   C1 × 6 × 10−6 12 × 6 Ceq = = × 10−6 = 4 μF. C1 + 6 × 10−6 12 + 6

Figure 5-17: Capacitors in parallel.

Capacitors in parallel

The equivalent circuit is shown Fig. 5-18(b).

The three capacitors shown in Fig. 5-17 share the same voltage υs and are therefore connected in parallel. The source current is is equal to the sum of their currents, is = i1 + i2 + i3 = C1

dυs dυs dυs + C2 + C3 . dt dt dt

1

(5.38)

R2 = 3 kΩ

For the equivalent circuit with equivalent capacitor Ceq , is = Ceq

dυs . dt

(5.39)

Equating the expressions given by Eqs. (5.38) and (5.39) leads to Ceq = C1 + C2 + C3 ,

R1 = 8 kΩ C1 = 12 μF

C2 = 1 μF

R3 = 6 kΩ

C3 = 5 μF

2

(5.40)

(a) Original circuit

which can be generalized to N capacitors in parallel as

Ceq =

N 

1 Ci

(capacitors in parallel).

(5.41)

i=1

Req = 10 kΩ Ceq = 4 μF

2

Since the capacitors are connected in parallel, they shared the same voltage υ(t0 ) at reference time t0 . Hence, for the equivalent capacitor υeq (t0 ) = υ(t0 ).

(b) Equivalent circuit Figure 5-18: Circuit for Example 5-5.

(5.42)

“book” — 2015/5/4 — 7:14 — page 265 — #18

TECHNOLOGY BRIEF 12: SUPERCAPACITORS

265

Technology Brief 12 Supercapacitors As shown in Section 5-2.1, the energy (in joules) stored in a capacitor is given by w = 12 CV 2 , where C is the capacitance and V is the voltage across it. Why then do we not charge capacitors by applying a voltage across them and then use them instead of batteries in support of everyday gadgets and systems? To help answer this question, we refer the reader to Fig. TF12-1, whose axes represent two critical attributes of storage devices. It is the combination (intersection) of these attributes that determines the type of applications best suited for each of the various energy devices displayed in the figure.

Energy density W  is a measure of how much energy a device or material can store per unit weight. That is, W  = w/m, where m is the mass of the capacitor in kilograms. [Alternatively, energy density can be defined in terms of volume (instead of weight) for applications where minimizing the volume of the energy source is more important than minimizing its weight.] Even though the formal SI unit for energy density is (J/kg), a more common unit is the watt-hour/kg (Wh/kg) with 1 Wh = 3600 J. The second dimension in Fig. TF12-1 is the power density P  (W/kg), which is a measure of how fast energy can be added to or removed from an energy-storage device (also per unit weight). Power is defined as energy per unit time as P  = dW  /dt.

Energy density W ’ (W-h/kg)

Charge/discharge time

Power density P ’ (W/kg)

Figure TF12-1: Energy and power densities of modern energy-storage technologies. Even though supercapacitors store less charge than batteries, they can discharge their energy more quickly, making them more suitable for hybrid cars. (Science, Vol. 313, p. 902.)

“book” — 2015/5/4 — 7:14 — page 266 — #19

266

TECHNOLOGY BRIEF 12: SUPERCAPACITORS

Table TT12-1: Comparison of a conventional capacitor, supercapacitor, and lithium battery size and mass required to hold ∼ 1 megajoule (MJ) of energy (300 watt-hours). 1 MJ of energy will power a laptop with an average consumption of 50 W for 6 hours. Note from the first column that a lithium ion battery might hold 1000 times more energy than a conventional capacitor for reasonable voltages (< 50 V).

Sample device Convenonal capacitor Supercapacitor Lithium ion baery

Specific Energy [Wa hours/ kg]

Specific Energy [MJ / kg]

Energy Density [MJ / liter]

Volume required to hold 1 MJ [liter]

Weight required to hold 1 MJ [kg]

0.01 – 0.1

4x10-5-4x10-4

6x10-5-6x10-4

17000-1700

25000 - 2500

1 - 10 100 - 250

0.004 – 0.04 0.36 - 0.9

0.006 - 0.06 1-2

166 – 16 1 – 0.5

250 – 25 2.8 – 1.1

According to Fig. TF12-1, fuel cells can store large amounts of energy, but they can deliver that energy only relatively slowly (several hours). In contrast, conventional capacitors can store only small amounts of energy— several orders of magnitude less than fuel cells—but it is possible to charge or discharge a capacitor in just a few seconds—or even a fraction of a second. Batteries occupy the region in-between fuel cells and conventional capacitors; they can store more energy per unit weight than the ordinary capacitor by about three orders of magnitude, and they can release their energy faster than fuel cells by about a factor of 10. Thus, capacitors are partly superior to other energy devices because they can accomodate very fast rates of energy transfer, but the amount of energy that can be “packed into” a capacitor is limited by its size and weight. To appreciate what that means, let us examine the relation w=

value of ε is in the range between ε0 (permittivity of vacuum) and 6ε0 (for mica), so the choice of material can at best increase C by a factor of 6. Making A larger increases both the volume and weight of the capacitor. In fact, since the mass m of the plates is proportional directly to A, the energy density W  = w/m is independent of A. That leaves d as the only remaining variable. Reducing d will indeed increase C, but such a course will run into two serious obstacles: (a) to avoid voltage breakdown (arcing), V has to be reduced along with d such that V/d remains lower than the breakdown value of the insulator; (b) eventually d approaches subatomic dimensions, making it infeasible to construct such a capacitor. Increasing V also increases the energy stored (by V 2 ) but here, too, we run into problems with breakdown. Another serious limitation of the capacitor as an energy storage device is that its voltage does not remain constant as energy is transferred to and from it.

1 CV 2 . 2

To increase w, we need to increase either C or V. We can develop an intuitive feel for this if we compare how large a storage element would have to be to hold 1 MJ (∼ 300 watt-hours). From Table TT12-1, we can see that a conventional capacitor would have to be thousands of liters in size (and weigh thousands of kilograms), whereas a supercapacitor or a battery would be considerably smaller. For a parallel-plate capacitor, C = εA/d, where ε is the permittivity of the material between the plates, A is the area of each of the two plates, and d is the separation between them. The material between the plates should be a good insulator, and for most such insulators, the

Supercapacitor Technology A new generation of capacitor technologies, termed supercapacitors or ultracapacitors, is narrowing the gap between capacitors and batteries. These capacitors can have sufficiently high energy densities to approach within 10 percent of battery storage densities, and additional improvements may increase this even more. Importantly, supercapacitors can absorb or release energy much faster than a chemical battery of identical volume. This helps immensely during recharging. Moreover, most batteries can be recharged only a few hundred times before they are degraded completely; supercapacitors can be charged and discharged millions

“book” — 2015/5/4 — 7:14 — page 267 — #20

TECHNOLOGY BRIEF 12: SUPERCAPACITORS

Outer Helmholtz Plane (OHP)

Solvated ion and hydration (water) sheet

267

Activated carbon

Separator Electrodes

5-10 nm

(a)

(b)

Figure TF12-2: (a) Conceptual illustration of the water double layer at a charged metal surface; (b) conceptual illustration of an electrochemical capacitor.

of times before they wear out. Supercapacitors also have a much smaller environmental footprint than conventional chemical batteries, making them particularly attractive for green energy solutions.

History and Design Supercapacitors are a special class of capacitor known as an electrochemical capacitor. This should not be confused with the term electrolytic capacitor, which is a term applied to a specific variety of the conventional capacitor. Electrochemical capacitors work by making use of a special property of water solutions (and some polymers and gels). When a metal electrode is immersed in water and a potential is applied, the water molecules (and any dissolved ions) immediately align themselves to the charges present at the surface of the metal electrode, as illustrated in Fig. TF12-2(a). This rearrangement generates a thin layer of organized water molecules (and ions), called a double layer, that extends over the entire surface of the metal. The very high charge density, separated by a tiny distance on the order of a few nanometers, effectively looks

like a capacitor (and a very large one: capacitive densities on the order of ∼ 10 μF/cm2 are common for water solutions). This phenomenon has been known to physicists and chemists since the work of von Helmholtz in 1853, and later Guoy, Chapman, and Stern in the early 20th century. In order to make capacitors useful for commercial applications, several technological innovations were required. Principal among these were various methods for increasing the total surface area that forms the double layer. The first working capacitor based on the electrochemical double layer (patented by General Electric in 1957) used very porous conductive carbon. Modern electrochemical capacitors employ carbon aerogels, and more recently carbon nanotubes have been shown to effectively increase the total double layer area (Fig. TF12-2(b)). Supercapacitors are beginning to see commercial use in applications ranging from transportation to lowpower consumer electronics. Several bus lines around the world now run with buses powered with supercapacitors; train systems are also in development. Supercapacitors intended for small portable electronics (like your MP3 player) are in the pipeline as well!

“book” — 2015/5/4 — 7:14 — page 268 — #21

268

CHAPTER 5 which reduces to

Example 5-6: Voltage Division

Figure 5-19(a) contains two resistors R1 and R2 connected in series to a voltage source υs . In Chapter 2, we demonstrated that the voltage υs is divided among the two resistors and, for example, υ1 is given by   R1 υs . (5.43) υ1 = R1 + R 2 Derive the equivalent voltage-division equation for the series capacitors C1 and C2 in Fig. 5-19(b).Assume that the capacitors had no charge on them before they were connected to υs . Solution: From the standpoint of the source υs , it “sees” an equivalent, single capacitor C given by the series combination of C1 and C2 , namely C=

RC AND RL FIRST-ORDER CIRCUITS

C1 C2 . C1 + C 2

(5.44)

C1 υ1 = C2 υ2 .

(5.47)

Using υ2 = υs − υ1 in Eq. (5.47) leads to C1 υ1 = C2 (υs − υ1 ) or

 C2 υ1 = υs . (5.48) C1 + C 2 We note that in the voltage-division equation for resistors, υ1 is directly proportional to R1 , whereas in the capacitor case, υ1 is directly proportional to C2 (instead of to C1 ). Additionally, in view of the relationship given by Eq. (5.47), application of the basic definition for capacitance, namely C = q/υ, leads to 

q1 = q2 .

(5.49)

This result is exactly what one would expect when viewing the circuit from the perspective of the voltage source υs .

The voltage across C is υs . The law of conservation of energy requires that the energy that would be stored in the equivalent capacitor C be equal to the sum of the energies stored in C1 and C2 . Hence, application of Eq. (5.29) gives

Concept Question 5-11: Compare the voltage-division

1 1 1 Cυs2 = C1 υ12 + C2 υ22 . 2 2 2

Concept Question 5-12: Two capacitors are connected

(5.45)

Upon replacing C with the expression given by Eq. (5.44) and replacing the source voltage with υs = υ1 + υ2 , we have   C1 C2 1 1 1 (υ1 + υ2 )2 = C1 υ12 + C2 υ22 , (5.46) 2 C1 + C 2 2 2

equation for two capacitors in series with that for two resistors in series. Are they identical or different in form? (See )

in series between terminals (a, b) in a certain circuit with capacitor 1 next to terminal a and capacitor 2 next to terminal b. How does the magnitude and polarity of charge q1 on the plate (of capacitor 1) near terminal a compare with charge q2 on the plate (of capacitor 2) near terminal b? (See ) Exercise 5-9: Determine Ceq and υeq (0) at terminals

Voltage Division + υ1 _ υs

+ _

R1 R2 

(a) υ1 =  υ2 =

+

R1 R1 + R 2 R2 R1 + R 2

+ υ _2

υs

+ _

q1 −q1 C1 q2 C2 −q2 

 υs

(b) υ1 = 

 υs

υ1 _

υ2 =

C2 C1 + C 2 C1 C1 + C 2

+ _ υ2

(a, b) for the circuit in Fig. E5.9 given that C1 = 6 μF, C2 = 4 μF, C3 = 8 μF, and the initial voltages on the three capacitors are υ1 (0) = 5 V and υ2 (0) = υ3 (0) = 10 V, respectively.

a

+

υ1 C1

 υs 

_ C2

+ _ υ2

+ _υ3

C3

b

υs

Figure 5-19: Voltage-division rules for (a) in-series resistors and (b) in-series capacitors.

Figure E5.9 Answer: Ceq = 4 μF, υeq(0) = 15 V. (See

C3

)

“book” — 2015/5/4 — 7:14 — page 269 — #22

5-3

INDUCTORS

269

Area S

Exercise 5-10: Suppose the circuit of Fig. E5.9 is

connected to a dc voltage source V0 = 12 V. Assuming that the capacitors had no charge before they were connected to the voltage source, determine υ1 and υ2 given that C1 = 6 μF, C2 = 4 μF, and C3 = 8 μF. Answer: υ1 = 8 V, υ2 = 4 V. (See

C3

i



)

5-3 Inductors Any current-carrying conductor, whether straight or coiled, forms an inductor. A current produces a magnetic field, which stores energy that can be released later in the form of another current. Also, since every wire acts like an inductor, we have small amounts of stray inductance in every circuit. Fortunately, this can be ignored except at extremely high frequencies (microwave band). Inductors exhibit a number of useful properties, including magnetic coupling and electromagnetic induction. They are employed in microphones and loudspeakers, magnetic relays and sensors, theft detection devices, and motors and generators, and they provide wireless power transmission and data communication (albeit over relatively short distances).  Capacitors and inductors constitute a canonical pair of devices. Whereas a capacitor can store energy through the electric field induced by the voltage imposed across its terminals, an inductor can store magnetic energy through the magnetic field induced by the current flowing through its wires.  The i–υ relationship for a capacitor is i = C dυ/dt; the converse is true for an inductor with υ = L di/dt. As we will see in Chapter 7, the capacitor acts like an open circuit to lowfrequency signals and like a short circuit to high-frequency signals; the exact opposite behavior is exhibited by the inductor. A typical example of an inductor is the solenoid configuration shown in Fig. 5-20. The solenoid consists of multiple turns of wire wound in a helical geometry around a cylindrical core. The core may be air filled or may contain a magnetic material (typically iron) with magnetic permeability μ. If the wire carries a current i(t) and the turns are closely spaced, the solenoid produces a relatively uniform magnetic field B within its interior region. Magnetic-flux linkage is defined as the total magnetic flux linking (passing through) a coil or a given circuit. For a solenoid with N turns carrying a current i,   μN 2 S i (Wb), (5.50) =

Core

Magnetic-field lines

Figure 5-20: The inductance of a solenoid of length and cross-sectional area S is L = μN 2 S/ , where N is the number of turns and μ is the magnetic permeability of the core material.

where is the length of the solenoid and S is its cross-sectional area. The unit for is the weber (Wb), named after the German scientist Wilhelm Weber (1804–1891). Self-inductance refers to the magnetic-flux linkage of a coil (or circuit) with itself, in contrast with mutual inductance, which refers to magnetic-flux linkage in a coil due to the magnetic field generated by another coil (or circuit). Usually, when the term inductance is used, the intended reference is to self-inductance. Mutual inductance is covered in Chapter 11. The (self) inductance of any conducting system is defined as the ratio of to the current i responsible for generating it, given as (H), (5.51) L= i and its unit is the henry (H), so named to honor the American inventor Joseph Henry (1797–1878). Using the expression for given by Eq. (5.50), we have

L=

μN 2 S

(solenoid).

(5.52)

The inductance L is directly proportional to μ, the magnetic permeability of the core material. The relative magnetic permeability μr is defined as μr =

μ , μ0

(5.53)

where μ0 ≈ 4 π × 10−7 (H/m) is the magnetic permeability of free space.

“book” — 2015/5/4 — 7:14 — page 270 — #23

270

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

Table 5-3: Relative magnetic permeability of materials, μr = μ/μ0 and μ0 = 4π × 10−7 H/m. Material All Dielectrics and Non-Ferromagnetic Metals Ferromagnetic Metals Cobalt Nickel Mild steel Iron (pure) Silicon iron Mumetal Purified iron

Relative Permeability μr

≈ 1.0 250 600 2,000 4,000–5,000 7,000 ∼ 100, 000 ∼ 200, 000

High current inductor

Planar inductor

Solenoid Figure 5-21: Various types of inductors.  Except for ferromagnetic materials, μr ≈ 1 for all dielectrics and conductors. According to Table 5-3, μr of ferromagnetic materials (which include iron, nickel, and cobalt) can be as much as five orders of magnitude larger than that of other materials. Consequently, L of an ironcore solenoid is about 5000 times that of an air-core solenoid of the same size and shape.  Air-core inductors have relatively low inductances, on the order of 10 μH or smaller. Consequently, they are used mostly in high-frequency circuits, such as those designed to support AM and FM radio, cell phones, TV, and similar types of transmitters and receivers. Ferrite-core inductors have the inductance-size advantage over air-core inductors, but they have the disadvantage that the ferrite material is subject to hysteresis effects, and they tend to be larger and heavier than their air-core counterparts. One of the consequences of magnetic hysteresis is that the inductance L becomes a function of the current flowing through it. Magnetic hysteresis is outside the scope of this book; hence, we will always assume that an inductor is an ideal linear device and its inductance is constant and independent of the current flowing through it. In modern circuit design and manufacturing, it is highly desirable to contain circuit size down to the smallest dimensions possible. To that end, it is advantageous to use planar integratedcircuit (IC) devices whenever possible. It is relatively easy to manufacture resistors and capacitors in a planar IC format and to do so for a wide range of resistance and capacitance values, but the same is not true for inductors. even though inductors can be manufactured in planar form, as illustrated by the coil shown in Fig. 5-21, their inductance values are too small for most circuit

applications, necessitating the use of the more bulky, discrete form instead.

5-3.1

Electrical Properties

According to Faraday’s law, if the magnetic-flux linkage in an inductor (or circuit) changes with time, it induces a voltage υ across the inductor’s terminals given by d . dt

(5.54)

di d (Li) = L . dt dt

(5.55)

υ= In view of Eq. (5.51), υ=

This i–υ relationship adheres to the passive sign convention introduced earlier for resistors and capacitors. If the direction of i is into the (+) voltage terminal of the inductor (Fig. 5-22), then the inductor is receiving power. Also, the same logic that led us earlier to the conclusion that the voltage across a capacitor cannot change instantaneously leads us now to the conclusion:  The current through an inductor cannot change instantaneously, but the voltage can.  (Otherwise, the voltage across it would become infinite.) The implication of this restriction is that when a current source connected to an inductor is disconnected by a switch, the current

“book” — 2015/5/4 — 7:14 — page 271 — #24

5-3

INDUCTORS

271 where it is presumed that at t = −∞ no current was flowing through the inductor. Note the analogy with the capacitor for which w(t) = 21 C υ 2 (t).

i L

+ υ _

υ=L

di dt

 The magnetic energy stored in an inductor at a given instant in time depends on the current flowing through the inductor at that instant—without regard to prior history. 

Figure 5-22: Passive sign convention for an inductor.

continues to flow for a short amount of time through the air between the switch terminals, manifesting itself in the form of a spark! In large power systems, current must always be ramped up and down slowly to avoid this problem. When we discussed the capacitor’s i–υ relationship given by Eq. (5.23), we noted that under dc conditions a capacitor acts like an open circuit. In contrast, Eq. (5.55) asserts that:  Under dc conditions, an inductor acts like a short circuit.  To express i(t) in terms of υ(t), we duplicate the procedure we followed earlier in connection with the capacitor, which for the inductor leads to

i(t) = i(t0 ) +

1 L

t

υ dt  ,

(5.56)

where t0 is an initial reference point in time. The power delivered to the inductor is given by



t 

p dt =

w(t) = −∞

−∞

di , dt

(5.57)

di Li  dt





dt ,

(5.58)

1 L i 2 (t) 2

(for t ≥ 0).

(a) Plot the waveform i(t) versus t and determine the locations of its first maximum, first minimum, and their corresponding amplitudes. (b) given that L = 50 mH, obtain an expression for υ(t) across the inductor and plot its waveform.

(J),

Solution: (a) The waveform of i(t) is shown in Fig. 5-23(b). To determine the locations of its maxima and minima, we take the derivative of i(t) and equate it to zero, which leads to π  2

× 10e−0.8t cos



πt 2

 = 0,

which in turn simplifies to  tan

πt 2

 =

π . 1.6

Its solution is πt = 1.1 + nπ 2

(for n = 0, 1, 2, . . . ).

For n = 0, t = 0.7 s, which is the location in time of the first maximum of i(t). The next solution, corresponding to n = 1, gives the location of the first minimum of i(t) at 2.7 s. The amplitudes of i(t) at these locations are

which yields w(t) =

i(t) = 10e−0.8t sin(π t/2) A,

−0.8 × 10e−0.8t sin(π t/2) +

and as with the resistor and the capacitor, the sign of p determines whether the inductor is receiving power (p > 0) or delivering it (p < 0). The accumulation of power over time constitutes the storage of energy. The magnetic energy stored in an inductor is t

Upon closing the switch at t = 0 in the circuit of Fig. 5-23(a), the voltage source generates a current waveform through the circuit given by

(c) Generate a plot of the power p(t) delivered to the inductor.

t0

p(t) = υi = Li

Example 5-7: Inductor Response to Current Waveform

(5.59)

imax = i(t = 0.7 s) = 10e−0.8×0.7 sin(π × 0.7/2) = 5.09 A

“book” — 2015/5/4 — 7:14 — page 272 — #25

272

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

i (A) R υs(t)

6 5.09

i(t)

4

+

t=0

+ _

L

Current

υ(t)

2

_

0 0 −1.03 −2

2

4

(a)

8

10

t (s)

(b) p (W)

υ (V)

Power

1.5

1 0.78

1

Power transfer into inductor (current increasing)

Voltage

0.5

0

6

0.5

0

2

4

6

8

10

t (s)

−0.5

0

0

2

−0.5

4 6 8 10 Power transfer out of inductor (current decreasing)

t (s)

−1 (c)

(d) Figure 5-23: Circuit for Example 5-7.

(c)

and imin = i(t = 2.7 s) = 10e−0.8×2.7 sin(π × 2.7/2) = −1.03 A.

p(t) = υ(t) i(t) = [−0.4 sin(π t/2) + 0.25π cos(π t/2)]e−0.8t × 10e−0.8t sin(π t/2)

(b) di υ(t) = L dt d = L [10e−0.8t sin(π t/2)] dt = 50 × 10−3 · [−8e−0.8t sin(π t/2) + 5πe−0.8t cos(π t/2)] = [−0.4 sin(π t/2) + 0.25π cos(π t/2)]e−0.8t V. The waveform of υ(t) is shown in Fig. 5-23(c).

= [−4 sin2 (π t/2) + 2.5π cos(π t/2) sin(π t/2)] × e−1.6t W. The waveform of p(t) shown in Fig. 5-23(d) includes both positive and negative values. During periods when p(t) > 0, magnetic energy is getting stored in the inductor. Conversely, when p(t) < 0, the inductor is releasing some of its previously stored energy. Concept Question 5-13: What type of material exhibits

a magnetic permeability higher than μ0? (See

)

“book” — 2015/5/4 — 7:14 — page 273 — #26

5-3

INDUCTORS

273

Concept Question 5-14: Can the voltage across an

inductor change instantaneously? (See

Combining In-Series Inductors

)

is

Exercise 5-11: Calculate the inductance of a 20-turn air-

core solenoid if its length is 4 cm and the radius of its circular cross section is 0.5 cm. Answer: L = 9.87 × 10−7 H = 0.987 μH. (See

C3

1

+ _

υs

+ υ1 _ + υ2 _ + υ3 _ L1

L2

L3

2

)

Exercise 5-12: Determine currents i1 and i2 in the circuit

of Fig. E5.12, under dc conditions. Answer: i1 = 0, i2 = 6 A. (See

is C

)

+ _

υs i1

L1 6A

L2

4 kΩ

1 Leq = L1 + L2 + L3 2

i2

Figure 5-24: Inductors in series.

L3

6 kΩ

Combining In-Parallel Inductors

Figure E5.12

5-3.2

is

Series and Parallel Combinations of Inductors

 The rules for combining multiple inductors in series or in parallel are the same as those for resistors. 

υs

1

+ _

i1

i2

i3

L1

L2

L3

2

Inductors in series For the three inductors in series in Fig. 5-24, dis dis dis υs = υ1 + υ2 + υ3 = L1 + L2 + L3 dt dt dt dis = (L1 + L2 + L3 ) , (5.60) dt and for the equivalent circuit, dis υs = Leq . (5.61) dt Hence, Leq = L1 + L2 + L3 , (5.62) and for N inductors in series, Leq =

N 

υs

1 

+ _

Leq =

2

 1 1 1 −1 + + L1 L2 L3

Figure 5-25: Inductors in parallel.

Inductors in parallel A similar analysis for the currents in the parallel circuit of Fig. 5-25 leads to

Li = L1 + L2 + · · · + LN

i=1

(inductors in series).

is

(5.63)

1 1 1 1 = + + . Leq L1 L2 L3

(5.64)

“book” — 2015/5/4 — 7:14 — page 274 — #27

274

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

Table 5-4: Basic properties of R, L, and C. Property

R

i–υ relation

υ R

i=

υ-i relation

υ = iR

p (power transfer in)

p = i2R

w (stored energy)

0

i=

1 L

t

L

C

υ dt  + i(t0 )

i=C

t0

di υ=L dt

1 υ= C

di p = Li dt 1 2 w = Li 2

t

i dt  + υ(t0 )

t0

p = Cυ

1 Cυ 2 2 1 1 1 = + Ceq C1 C2

Leq = L1 + L2

1 1 1 = + Req R1 R2 no change

1 1 1 = + Leq R1 R2 short circuit

Can υ change instantaneously?

yes

yes

no

Can i change instantaneously?

yes

no

yes

Parallel combination dc behavior

dυ dt

w=

Req = R1 + R2

Series combination

dυ dt

Ceq = C1 + C2 open circuit

Generalizing to the case of N inductors,

L3 = 1 mH

2 kΩ 1 = Leq

N  i=1

1 1 1 1 = + + ··· + . Li L1 L2 LN

(5.65)

C1 = 10 μF

If i1 (t0 ) through iN (t0 ) are the initial currents flowing through the parallel inductors L1 to LN at t0 , then the initial current ieq (t0 ) that would be flowing through the equivalent inductor Leq is given by

ieq (t0 ) =

N 

ij (t0 ).

+ _

C2 = 4 μF 24 V

(a) Original circuit

2 kΩ

(5.66)

I1

L3

V

6 kΩ

I2

j =1

A summary of the electrical properties of resistors, inductors and capacitors is available in Table 5-4.

4 kΩ

L2 = 0.5 mH

L1 = 0.2 mH

(inductors in parallel)

6 kΩ

L2 L1 C1

4 kΩ + _

C2 24 V

Example 5-8: Energy Storage under dc Conditions

(b) Equivalent circuit under steady state conditions The circuit in Fig. 5-26(a) has been in its present state for a long time. Determine the amount of energy stored in the capacitors and inductors.

Figure 5-26: Under steady-state dc conditions, capacitors act like open circuits, and inductors act like short circuits.

“book” — 2015/5/4 — 7:14 — page 275 — #28

5-4

RESPONSE OF THE RC CIRCUIT

275

Solution: Our first step is to replace components with their dc equivalents (capacitors with open circuits and inductors with short circuits). The process leads to the circuit in Fig. 5-26(b), which can be solved using any of the analysis methods used previously with resistive circuits. Current I1 then is given by 24 = 4 mA, I1 = (2 + 4)k and node voltage V is V = 24 − (4 × 10−3 × 4 × 103 ) = 8 V. Hence, the amounts of energy stored in C1 , C2 , L1 , L2 , and L3 are 1 1 C 1 : W = C1 V 2 = × 10−5 × 64 = 0.32 mJ, 2 2 1 1 C 2 : W = C2 V 2 = × 4 × 10−6 × 64 = 0.128 mJ, 2 2 1 L1 : W = L1 I12 2 1 = × 0.2 × 10−3 × (4 × 10−3 )2 = 1.6 nJ, 2 1 1 L2 : W = L2 I22 = × 0.5 × 10−3 × (0) = 0, 2 2 and 1 1 L3 : W = L3 I12 = × 10−3 × (4 × 10−3 )2 = 8 nJ. 2 2 Concept Question 5-15: How do the rules for adding inductors in series and in parallel compare with those for resistors and capacitors? (See ) Concept Question 5-16: An inductor stores energy through the magnetic field B, but the equation for the energy stored in an inductor is w = 1 Li2. Explain. 2 (See )

Exercise 5-13: Determine Leq at terminals (a, b) in the circuit of Fig. E5.13.

a

5-4

Response of the RC Circuit

The preceding sections described the behavior of capacitors and inductors under dc conditions (i.e., a static circuit with none of its voltages or currents varying with time). We now turn our attention to the time-varying (dynamic) conditions of these circuits. From the standpoint of analysis and design, circuits containing capacitors and inductors are divided into three groups: • RC Circuits: composed of sources (either constant or time-varying), capacitors, and resistors. • RL Circuits: composed of sources (either constant or time-varying), inductors, and resistors. • RLC Circuits: composed of any combination and any number of sources, capacitors, inductors, and resistors. In this and succeeding sections of this chapter, we examine the responses of relatively simple RC and RL circuits to sudden changes, such as closing or opening a switch—or both sequentially—and we limit the sources to dc voltage and current sources. The RLC circuit response is addressed in Chapter 6, also for dc sources with switches. RLC circuits driven by ac sources are treated in Chapters 7–11, and RLC circuits driven by other types of sources are the subject of Chapters 12 and 13. The circuit shown in Fig. 5-27 is called a first-order RC circuit; it contains a resistor and a capacitor, and its current and voltage responses are determined by solving a first-order differential equation. The name also applies to any other circuit containing sources, resistors, and capacitors—provided it can be reduced to the form of the generic RC circuit of Fig. 5-27 or its Norton equivalent. This can be realized by combining elements in series or in parallel, as well as through Y- transformations. The voltage source exciting the circuit is a rectangular pulse of amplitude Vs and duration T0 , which includes both turn-on (charging) and turn-off (discharging) periods. The objective of the present section is to develop a

2 mH

R

6 mH

Vs

12 mH

b

+ _

υi = t=0 Figure E5.13

Answer: Leq = 6 mH. (See

)

iC

C

t = T0

Figure 5-27: Generic first-order RC circuit.

+ υ _ C

“book” — 2015/5/4 — 7:14 — page 276 — #29

276

CHAPTER 5

methodology appropriate for RC circuits, so we may apply it to evaluate the circuit’s response to the rectangular-pulse waveform or to other types of nonperiodic waveforms.

5-4.1

Natural Response of a Charged Capacitor

We begin by considering what is called the natural response of the circuit, which refers to the time variations of the voltages and currents in reaction to moving a switch that allows a fully charged capacitor to discharge its accumulated charge. This occurs at t = T0 in Fig. 5-27. To that end, let us examine the more realistic circuit in Fig. 5-28(a). Until t = 0, the series RC circuit had been connected to dc voltage source Vs for a long time. At t = 0, the switch disconnects the RC circuit from the

Rs

Vs

R

1



iC = C dtC

t=0

+ _

2

C

+ _ υC

Vs

R

1

iC = 0

+ − C _υC(0 )

+ _

= Vs

(b) At t =

0−

source and connects it to terminal 2. We seek to determine the voltage response of the capacitor υ(t) for t ≥ 0. Before we start our solution, it is important to consider the implication of the information we are given about the state of the capacitor before and after moving the switch. For purposes of clarity, we define: (a) t = 0− as the instant just before the switch is moved from terminal 1 to terminal 2, and (b) t = 0 as the instant just after it was moved; t = 0 is synonymous with t = 0+ . At t = 0− , the circuit had been in the condition shown in Fig. 5-28(a) for a long time. As we noted earlier in Section 5-2.1, when a dc circuit is in a steady state, its capacitors act like open circuits. Consequently, the open circuit in Fig. 5-28(b), representing the state of the circuit at t = 0− , allows no current to flow through the loop, and, therefore, there is no voltage drop across either of the two resistors. Hence, υC (0− ) = Vs , and since the voltage across the capacitor cannot change instantaneously, it follows that υC (0), the voltage after moving the switch, is given by υC (0) = υC (0− ) = Vs .

RiC + υC = 0

RC

2



C

(c) At t > 0 (capacitor discharging) Figure 5-28: RC circuit with an initially charged capacitor that starts to discharge its energy after t = 0.

(5.68)

dυC + υC = 0. dt

(5.69)

Upon dividing both terms by RC, Eq. (5.69) takes the form

iC = C dtC + _ υC

(for t ≥ 0),

where iC is the current through and υC is the voltage across the capacitor. Since iC = C dυC /dt, Eq. (5.68) becomes

(fully charged capacitor)

R

(5.67)

As we see shortly, we will need this piece of information for when we apply this initial condition to the solution of the differential equation of υC (t). For t ≥ 0, application of KVL to the loop in Fig. 5-28(c) gives

(a) RC circuit Rs

RC AND RL FIRST-ORDER CIRCUITS

dυC + aυC = 0 dt

(source-free),

(5.70)

where 1 . (5.71) RC When arranging a differential equation in υC (t), it is customary to place all terms that involve υC (t) on the left-hand side of the equation and to place terms that do not involve υC (t) on the right-hand side. The term(s) on the right-hand side is (are) called a=

“book” — 2015/5/4 — 7:14 — page 277 — #30

5-4

RESPONSE OF THE RC CIRCUIT

277

the forcing function. For a circuit, the forcing function is related directly to the voltage and current sources in the circuit. Because the RC circuit in Fig. 5-28(c) does not contain any sources, Eq. (5.70) has a zero on its right-hand side and it is called (appropriately) a source-free, first-order differential equation.  The solution of the source-free equation is called the natural response (discharging condition) of the circuit.  The standard procedure for solving Eq. (5.70) starts by replacing t with dummy variable t  and multiplying both sides  by eat , dυC at   e + aυC eat = 0. dt 

(5.72)

Next, we recognize that the sum of the two terms on the left hand side is equal to the expansion of the differential of (υC eat ), dυC at  d   (υC eat ) = e + aυC eat .   dt dt

(5.73)

Hence, Eq. (5.72) becomes d  (υC eat ) = 0. dt 

0

d  (υC eat ) dt  = 0,  dt

υC (t) = υC (0) e−t/τ ,

(5.75)

where we have chosen the lower limit to be t  = 0 (because we are given specific information on the state of the circuit at that point in time). Performing the integration gives   t υC eat  = 0

(5.78)

(natural response discharging), with τ = RC

(s),

(5.79)

where τ is called the time constant of the circuit, and it is measured in seconds (s). In view of the initial condition given by Eq. (5.67), namely υC (0) = Vs , the expression for υC (t) becomes υC (t) = Vs e−t/τ u(t),

(5.80)

where we inserted the unit step function u(t) as a multiplication factor as a substitute for “for t ≥ 0.” The plot shown in Fig. 5-29(a) indicates that in response to the switch action, υC (t) decays exponentially with time from Vs at t = 0 down to its final value of zero as t → ∞. The decay rate is dictated by the time constant τ . At t = τ , υC (t = τ ) = Vs e−1 = 0.37Vs ,

(5.74)

Integrating both sides, we have t

The coefficient of t in the exponent is a critically important parameter, because it determines the temporal rate of υC (t). It is customary to rewrite Eq. (5.77) in the form

(5.81)

which means that at τ seconds after activating the switch, the voltage across the capacitor is down to 37 percent of its initial value. At t = 2τ , it reaches 14 percent, and at t = 5τ , it is less than 1 percent of its initial value. Hence, for all practical purposes, we can treat the circuit as having reached its final state when the switch has been in its new configuration for a time equal to or longer than 5τ .  The magnitude of the time constant τ is a measure of how fast or how slowly a circuit responds to a sudden change. 

0

or υC (t) eat − υC (0) = 0.

(5.76)

Solving for υC (t), we have υC (t) = υC (0) e−at = υC (0) e−t/RC

(for t ≥ 0), (5.77)

where we used Eq. (5.71) for a and appended the inequality t ≥ 0 to indicate that the expression given by Eq. (5.77) is valid only for t ≥ 0.

As we will see later in Section 5-7, the clock speed of a computer processor is, to first order, proportional to 1/τ . Hence, a slow circuit with τ = 1 ms would have a clock speed on the order of 1 kHz, whereas a fast circuit with τ = 1 ns can support clock speeds as high as 1 GHz. The current iC (t) flowing through the capacitor is given by iC (t) = C

dυC d =C (Vs e−t/τ ) dt dt Vs −t/τ = −C e τ

(for t ≥ 0),

(5.82)

“book” — 2015/5/4 — 7:14 — page 278 — #31

278

CHAPTER 5

υC Vs

Voltage discharging υC(t) = Vse−t/τ

0.37Vs t

τ

0

(a) iC τ −0.37 −

Vs R

t

0 iC(t) = −

Vs R

Current

Vs −t/τ e R

(b) pC τ/2 −0.37 −

Vs2

t

0

R

pC(t) = −

Vs2 R

Power

Vs2 −2t/τ e R

CVs2 0.37 2

( )

Vs −t/τ V2 × Vs e−t/τ = − s e−2t/τ u(t). e R R (5.84) Note that from the definition of u(t) given by Eq. (5.2), u(t) · u(t) = u(t). In general, power transfer is into a device if pC > 0 and out of it if pC < 0. Prior to t = 0, the capacitor had been connected to the voltage source for a long time. Hence, power already had flowed into the capacitor and was stored as electrical energy. The minus sign in Eq. (5.84) denotes that after t = 0 power flows out of the capacitor and gets dissipated in the resistor. pC (t) = iC υC = −

The amount of energy wC (t) contained in the medium between the capacitor’s oppositely charged conducting plates can be calculated either by integrating pC (t) over time from 0 to t or by applying Eq. (5.29). The latter approach gives

wC CVs2 wC(t) =

Fig. 5-29(b) indicates that after closing the switch at t = 0, the current changes instantly to (−Vs /R)—as if the capacitor were a voltage source Vs —and then it decays exponentially down to zero. The negative sign of i signifies that it flows in a counterclockwise direction through the loop, consistent with the behavior of the capacitor as a voltage source. Given υC (t) and iC (t), we can provide an expression for pC (t), the instantaneous power getting transferred to the capacitor, as

 The decay rate for pC (t) is 2/τ , which is twice as fast as that for υC (t) or iC (t). 

(c) 1 2

RC AND RL FIRST-ORDER CIRCUITS

1 2 −2t/τ 2 CVs e

wC (t) =

Energy t

1 CVs2 −2t/τ u(t). C υC2 (t) = e 2 2

(5.85)

(d)

Parts (c) and (d) of Fig. 5-29 display the time waveforms of pC (t) and wC (t), respectively.

Figure 5-29: Response of the RC circuit in Fig. 5-28(a) to

Concept Question 5-17: What specific characteristic

0

τ/2

defines a first-order circuit? (See

moving the SPDT switch to terminal 2.

Concept Question 5-18: What does the time constant

which simplifies to Vs −t/τ e u(t), R (natural response discharging) iC (t) = −

)

(5.83)

where, again, u(t) is used to emphasize the fact that the expression is valid for only t ≥ 0. The plot of iC (t) shown in

of an RC circuit represent? Would a larger capacitor discharge faster or more slowly than a small one? (See ) Concept Question 5-19: For the natural response of an

RC circuit, how does the decay rate for voltage compare with that for power? (See )

“book” — 2015/5/4 — 7:14 — page 279 — #32

5-4

RESPONSE OF THE RC CIRCUIT

279

Exercise 5-14: If in the circuit of Fig. E5.14

t=0 20 kΩ

Vs1

+ _υC

5 μF

R

1

υC (0− ) = 24 V, determine υC (t) for t ≥ 0.

iC

t=0

+ _

2

C

+ _

Vs2

+ _υC

(a) RC circuit Figure E5.14 Answer: υC

5-4.2

(t) = 24e−10t V

for t ≥ 0. (See

Rs C3

R

1

iC = 0

)

General Form of the Step Response of the RC Circuit

When we use the term circuit response, we mean the reaction of a certain voltage or current in the circuit to change, such as the introduction of a new source, the elimination of a source, or some other change in the circuit configuration. Whenever possible, we usually designate t = 0 as the instant at which the change occurred and t ≥ 0 as the time interval over which we seek the circuit response. In the general case, the capacitor may start with a voltage υC (0) at t = 0 (immediately after the sudden change) and may approach a value denoted υC (∞) as t → ∞. A circuit configuration that can represent such a scenario is the series RC circuit shown in Fig. 5-30(a). Prior to t = 0, the RC circuit is connected to a source Vs1 , and after t = 0, it is connected to a different source Vs2 . The circuit can be reduced to the following special cases:

Vs1

+ _

C

+ − _υC(0 ) = Vs 1

(b) Initial condition at t = 0−

iC

R

2 Vs2

C

+ _

+ _υC

(c) Natural reponse after t = 0 Figure 5-30: RC circuit switched from source Vs1 to source Vs2 at t = 0.

• Step response (due to Vs2 ) of an uncharged capacitor (if Vs1 = 0) • Step response (due to Vs2 ) of a charged capacitor (if Vs1 = 0)

For t ≥ 0, the (natural response) voltage equation for the loop in Fig. 5-30(c) is

• Natural response (if Vs2 = 0) of a charged capacitor (Vs1 = 0)

−Vs2 + iC R + υC = 0.

For obvious reasons, we excluded the trivial case where both Vs1 and Vs2 are zero, and we will now treat the general case where neither Vs1 nor Vs2 is zero. At t = 0− (Fig. 5-30(b)), the capacitor has been in steady state for a long time. Hence, it acts like an open circuit. Consequently, iC (0− ) = 0, and υC (0− ) = Vs1 . Since υC across the capacitor cannot change in zero time, the (initial condition) voltage υC (0) after moving the switch to terminal 2 is υC (0) = υC (0− ) = Vs1 .

(5.86)

(5.87)

Upon using iC = C dυC /dt and rearranging its terms, Eq. (5.87) can be written in the differential-equation form dυC + aυC = b, dt

(5.88)

where a=

1 RC

and

b=

V s2 . RC

(5.89)

“book” — 2015/5/4 — 7:14 — page 280 — #33

280

CHAPTER 5

We note that Eq. (5.88) is similar to Eq. (5.70), except that now we have a non-zero term on the right-hand side of the equation. Nevertheless, the method of solution remains the same. After replacing t with dummy variable t  and multiplying both sides  of Eq. (5.88) by eat , we have e

at 

dυC   + aυC eat = beat . dt 

(5.90a)

(5.90b)

0

d  (υC eat ) dt  = dt 

gives  υC eat |t0

t be

at 

dt



If the switch action causing the change in voltage across the capacitor occurs at time T0 instead of at t = 0, Eq. (5.96) assumes the form

· u(t − T0 ),

(5.98)

(5.91)

(5.92)

Series RC Circuit Solution

b at b e − , a a

(5.93)

1: If switch action is at t = 0, analyze circuit at t = 0− to determine initial conditions υC (0− ) and iC (0− ). Use this information to determine υC (0) and iC (0), at t immediately after the switch action. Remember that the voltage across a capacitor cannot change instantaneously (between t = 0− and t = 0), but the current can.

b (1 − e−at ). a

(5.94)

2: Analyze the circuit to determine υC (∞), the voltage across the capacitor long after the switch action.

Upon evaluating the functions at the two limits, we have

and then solving for υC (t), we have υC (t) = υC (0) e−at +

(5.97)

where we have replaced t with (t −T0 ) on the right-hand side of Eq. (5.96). Now υC (T0 ) is the initial voltage at t = T0 . For easy reference, this expression is made available in Table 5-5, along with expressions for three other types of circuits discussed in future sections.

0

 b at  t = e  . a 0

υC (t) eat − υC (0) =

υC (t) = Vs2 + (Vs1 − Vs2 )e−t/τ .

(series RC circuit with switch action at t = T0 )

Integrating both sides from t  = 0 to t  = t, namely t

For the specific circuit in Fig. 5-30(a), Eqs. (5.86) and (5.95) give υC (0) = Vs1 and υC (∞) = Vs2 . Hence,

  υC (t) = υC (∞) + [υC (T0 ) − υC (∞)]e−(t−T0 )/τ

In view of Eq. (5.73), Eq. (5.90a) can be rewritten as d   (υC eat ) = beat .  dt

RC AND RL FIRST-ORDER CIRCUITS

As t → ∞, e−∞ = 0 and υC (t) reduces to the final condition b υC (∞) = = Vs2 . a

(5.95)

By reintroducing the time constant τ = RC = 1/a and replacing b/a with υC (∞), we can rewrite Eq. (5.94) in the general form:   υC (t) = υC (∞) + [υC (0) − υC (∞)]e−t/τ u(t). (series RC circuit with switch action at t = 0) (5.96)  The voltage response of any RC circuit is determined by three parameters: the initial voltage υC (0), the final voltage υC (∞), and the time constant τ . 

3: Determine the time constant τ = RC. 4: Incorporate the information obtained in the previous three steps in Eq. (5.96):   υC (t) = υC (∞) + [υC (0) − υC (∞)]e−t/τ u(t). 5: If the switch action is at t = T0 instead of t = 0, replace 0 with T0 and use Eq. (5.98):   υC (t) = υC (∞) + [υC (T0 ) − υC (∞)] · e−(t−T0 )/τ · u(t − T0 ).

5-4.3 Th´evenin Approach For a circuit containing dc sources, resistors, switches and a single capacitor (or multiple capacitors that can be combined

“book” — 2015/5/4 — 7:14 — page 281 — #34

5-4

RESPONSE OF THE RC CIRCUIT

281

into a single equivalent capacitor), the voltage response across the capacitor, υC (t), can be calculated with relative ease by taking advantage of the Th´evenin theorem. The procedure involves the following steps:

a C

Subcircuit 1

+ υ _C

Subcircuit 2

b (a) Original circuit

Th´evenin Approach to RC Response Step 1: If the circuit includes a single switch action (open, close, or move between two terminals) at t = T0 , analyze the circuit at t = T0− (just before the switch action) to determine υC (T0− ). When so doing, the capacitor should be replaced with an open circuit. Then set υC (T0 ) = υC (T0− ), where υC (T0 ) is the voltage across the capacitor after the switch action. Step 2: For the circuit configuration at t ≥ T0 (after the switch action), obtain the Th´evenin equivalent circuit as “seen” by the capacitor. Figure 5-31(a) depicts a general circuit (composed of possibly two subcircuits) connected to a capacitor C. After removing (temporarily) the capacitor and calculating VTh and RTh of the equivalent Th´evenin circuit at terminals (a, b), reinstate the capacitor as in Fig. 5-31(b).

RTh VTh

+ _

a C

+ υ _C

b (b) After replacing circuit with Thévenin equivalent Figure 5-31: Replacing a resistive circuit with its Th´evenin equivalent as seen by capacitor C.

Step 3: The capacitor’s voltage response is then given by 

υC (t) = υC (∞) + [υC (T0 ) − υC (∞)]e

−(t−T0 )/τ



· u(t − T0 ), with υC (∞) = VTh , υC (T0 ) as obtained in step 1, and τ = RTh C. Step 4: If the circuit undergoes multiple switch actions, repeat the procedure for each time segment and use the property that the voltage across a capacitor cannot change instantaneously to match the responses at the boundaries between adjacent time segments.

of the circuit. Hence, no voltage drop occurs across the 3 k

resistor. Consequently, the voltage at node V1 , relative to the designated ground node, is V1 = 24 V. On the right-hand side of the circuit, the current source flows entirely through the 4 k resistor, generating a node voltage V2 = 4.5 × 10−3 × 4 × 103 = 18 V. Hence, the initial voltage is υC (0− ) = V1 − V2 = 24 − 18 = 6 V.

Example 5-9: Thevenin ´ Approach

The switch in the circuit of Fig. 5-32(a) had been in position 1 for a long time until it was moved to position 2 at t = 0. Determine υC (t) for t ≥ 0. Solution: Step 1: Figure 5-32(b) depicts the state of the circuit at t = 0− (initial condition), with the capacitor represented by an open circuit. Because of the open circuit, i = 0 in the left-hand side

Since the voltage across the capacitor cannot change instantaneously, it follows that υC (0) = υC (0− ) = 6 V. Step 2: Figure 5-32(c) represents the state of the circuit after moving the switch to position 2 and removing the capacitor so as to calculate the elements of the Th´evenin circuit at terminals (a, b). In step (d), conversion of the current source and 4 k

“book” — 2015/5/4 — 7:14 — page 282 — #35

282

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

Table 5-5: Response forms of basic first-order circuits. Diagram

Circuit

Response

R Input: dc circuit with switch action @ t = T0

RC

C

(τ = RC)

Input: dc circuit with switch action @ t = T0

RL

υC

  υC (t) = υC (∞) + [υC (T0 ) − υC (∞)]e−(t−T0 )/τ u(t − T0 )

iL R

L

  iL (t) = iL (∞) + [iL (T0 ) − iL (∞)]e−(t−T0 )/τ u(t − T0 ) (τ = L/R)

C R − + _ υi

Ideal integrator

υout

+

υout (t) = −

RL

1 RC

t

υi dt  + υout (t0 )

t0

R C − + _ υi

Ideal differentiator

+

υout RL

resistor into a voltage source in series with a resistor leads to RTh = 4 k + 1 k = 5 k , VTh = −4.5 × 10

−3

υout (t) = −RC

× 4 × 10 = −18 V.

Step 3: The capacitor is reinserted in part (e). With υC (0) = 6 V, υC (∞) = VTh = −18 V, and τ = RTh C = 5 × 103 × 100 × 10−6 = 0.5 s,

3

 Note that the polarity of the Th´evenin voltage source has to be assigned to match that of υC , the voltage across the capacitor. In the present case, the current to voltage transformation led to a voltage source with the opposite polarity to that defined for VTh . Hence, VTh = −18 V, not 18 V. 

dυi dt

we have

  υC (t) = υC (∞) + [υC (0) − υC (∞)]e−t/τ u(t) = [−18 + 24e−2t ] u(t) V.

This solution indicates that at t = 0, the initial voltage across the capacitor is υC (0) = −18 + 24 = 6 V, which is consistent with the result obtained in step 1. After a long time t such that e−2t approaches zero, υC (t) approaches −18 V, which is υC (∞). In between, the capacitor discharges to zero and

“book” — 2015/5/4 — 7:14 — page 283 — #36

5-4

RESPONSE OF THE RC CIRCUIT

+ υC _

3 kΩ t=0

1 24 V

+ _

283

V1

3 kΩ

4.5 mA

2

24 V

4 kΩ

+ _

2

4.5 mA

1 kΩ

1 kΩ

+ VTh _ a

+ VTh _

b

a

4 kΩ

4 kΩ

(b) Initial condition at t = 0−

(a) Circuit with switch

2

V2

a b C = 100 μF

1

i=0

C = 100 μF

υC(0−)_

+

b

2

4.5 mA

1 kΩ

RTh + _

υC

_

5 kΩ

C = 100 μF VTh = 18 V

4 kΩ

1 kΩ

(c) At t > 0 without the capacitor

+ _

18 V

+

(e) At t > 0, after reinserting C in the Thévenin equivalent circuit

(d) After current to voltage source transformation υC (V)

6

6 5 υC(t) = (−18 + 24e−2t ) u(t)

0 0

0.5

1

1.5

2

t (s)

−5 −10 −15 −18

−20 (f ) Plot

Figure 5-32: Circuit for Example 5-9. then builds up charge again, but of opposite polarity. The time variation of υC (t) is displayed in Fig. 5-32(f).

long time. Determine the voltage υC (t) for t ≥ 0 if the switch is moved at (a) t = 0 or (b) t = 3 s.

Example 5-10: Switching between Two Sources

Solution: (a) For T0 = 0 and t ≥ 0, the complete solution of υC (t) is given by Eq. (5.96) as   υC (t) = υC (∞) + [υC (0) − υC (∞)]e−t/τ u(t). (5.99)

In the circuit of Fig. 5-33(a), the SPDT switch is moved from position 1 to position 2 after it had been in position 1 for a

We need to determine three quantities: the initial voltage υC (0), the final voltage υC (∞), and the time constant τ .

“book” — 2015/5/4 — 7:14 — page 284 — #37

284

CHAPTER 5

i1

4 kΩ

45 V

1

2 i2

RC AND RL FIRST-ORDER CIRCUITS

2 kΩ

24 kΩ

t=0

+ _

8 kΩ 20 μF

+ _ υC

+ _

12 kΩ

60 V

(a) Original circuit i1 = 0

4 kΩ

45 V

+ _

1

2.67 kΩ

8 kΩ

30 V

+ _ _υC(0 )

C

C

20 μF

+ _υC

' Thevenin equivalent

2

24 kΩ + _

12 kΩ

60 V 20 μF

Circuit

+ _ _υC(0 )

= 30 V

(b) At t = 0− (initial condition) 2 kΩ

1

+ _

Circuit

2 i2

i1 = 0

i2

10 kΩ + _

+ _υC

20 V

' Thevenin equivalent

(c) At t ≥ 0 (steady state) Figure 5-33: Circuit for Example 5-10 [part (a)].

The initial voltage is the voltage that existed across the capacitor before moving the switch. Since the switch had been in that position for a long time, we presume that the circuit in Fig. 5-33(b) had reached its steady-state condition long before the switch was moved. Hence, at t = 0− (just before moving the switch), the capacitor behaves like an open circuit. The voltage υC (0− ) across the capacitor is the same as that across the 8 k resistor, and since i1 = 0 at t = 0− , application of voltage division yields υC (0− ) =



8k 4k + 8k

 × 45 = 30 V.

Incidentally, we could have obtained the same result by transforming the circuit in Fig. 5-33(b) into its Th´evenin equivalent. Incorporating the constraint that the voltage across the capacitor cannot change instantaneously, it follows that υC (0) = υC (0− ) = 30 V. Now we turn our attention to finding υC (∞). After moving the switch to position 2 (Fig. 5-33(c)) and allowing the circuit sufficient time to reach its final state, the capacitor again will

“book” — 2015/5/4 — 7:14 — page 285 — #38

5-4

RESPONSE OF THE RC CIRCUIT

285

behave like on open circuit, which means that i2 = 0 at t = ∞. Voltage division gives  υC (∞) =

12k 12k + 24k

 × 60 = 20 V.

The time constant of the circuit to the right of terminal 2 is given by τ = RC, with R being the Th´evenin resistance of that circuit. After suppressing (short-circuiting) the 60 V source, we get

12k × 24k = 10 k . 12k + 24k

R2 t = 10 s

+ _

V1

1

(a)

C

+ _ υC

C

+ _υ1

C

+ _υ2

Actual circuit R1

R = RTh = 2 k + 12 k  24 k

= 2 k +

2 t=0

R1

R2

2

+ _

V1

Hence,

Circuit during 0 ≤ t ≤ 10 s

(b) τ = RC = 10 × 10 × 20 × 10 3

−6

= 0.2 s.

R2

Substituting the values we obtained for υC (0), υC (∞), and τ in Eq. (5.99) leads to

Circuit after t = 10 s 1

υC (t) = [(20 + 10e−5t ) u(t)] V. (b) This is a repetition of the previous case except that now the switch action takes place at T0 = 3 s. The applicable expression is given by Eq. (5.98),

(c) υC (V) 7

  υC (t) = υC (∞) + [υC (3) − υC (∞)]e−(t−3)/τ u(t − 3).

6

Of course, υC (t) = 30 V before t = 3 s. Hence, for the specified time duration t ≥ 0,

4

2

Charging

Discharging

1 0

(d)

Given that the switch in Fig. 5-34 was moved to position 2 at t = 0 (after it had been in position 1 for a long time) and then returned to position 1 at t = 10 s, determine the voltage response υC (t) for t ≥ 0 and evaluate it for V1 = 20 V, R1 = 80 k , R2 = 20 k , and C = 0.25 mF.

υ2(t) = 6.59e−0.2(t − 10) V (for t ≥ 10 s)

3

30 V for 0 ≤ t ≤ 3 s, [20 + 10e−5(t−3) ] V for t ≥ 3 s.

Example 5-11: Charge/Discharge Action

υ2(t)

υ1(t)

5

 υC (t) =

υ1(t) = 20(1 − e−0.04t ) V (for 0 ≤ t ≤ 10 s)

6.59

0

5

10

15

20

25

30

t (s)

Voltage response

Figure 5-34: After having been in position 1 for a long time, the switch is moved to position 2 at t = 0 and then returned to position 1 at t = 10 s (Example 5-11). Solution: We will divide our solution into two time segments: υC = υ1 (t) for 0 ≤ t ≤ 10 s and υC = υ2 (t) for t ≥ 10 s.

“book” — 2015/5/4 — 7:14 — page 286 — #39

286

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

Time Segment 1: 0 ≤ t ≤ 10 s When the switch is in position 2 (Fig. 5-34(b)), the resistance of the circuit is R = R1 + R2 . Hence, the time constant during this first time segment is

Vs = 10 V

During 0 ≤ t ≤ 4 s Vs

+ _

C

+ _ υ1

C

+ _ υ2

(b) i

υ2 (t) = υ2 (∞) + [υ2 (10) − υ2 (∞)]e−(t−10)/τ2 . The new time constant is associated with the capacitor circuit after returning the switch to position 1,

The initial voltage υ2 (10) is equal to the capacitor voltage υ1 at the end of time segment 1, namely

(c) υC (V) Forced 6 response

2

With no voltage source present in the R2 C circuit, the charged capacitor will dissipate its energy into R2 , exhibiting a natural response with a final voltage of υ2 (∞) = 0. Consequently,

(d) (for t ≥ 10 s).

The complete time response of υ(t) is displayed in Fig. 5-34(d). Example 5-12: RC-Circuit Response to Rectangular Pulse

Determine the voltage response of a previously uncharged RC circuit to a rectangular pulse υi (t) of amplitude Vs and duration T0 , as depicted in Fig. 5-35(a). Evaluate and plot the response for R = 25 k , C = 0.2 mF, Vs = 10 V, and T0 = 4 s.

Natural response

4

υ2 (10) = υ1 (10) = 20(1 − e−0.04×10 ) = 6.59 V.

−(t−10)/τ2

R

After t = 4 s

τ2 = R2 C = 20 × 103 × 0.25 × 10−3 = 5 s.

= 6.59e−0.2(t−10) V

i

R

Voltage υ2 (t), corresponding to the second time segment (Fig. 5-34(c)), is given by Eq. (5.98) with a new time constant τ2 as

υ2 (t) = υ2 (10) e

+ _ υC

(a)

(for 0 ≤ t ≤ 10 s).

Time Segment 2: t ≥ 10 s

C

t=0 t=4s Pulse excitation

υ1 (t) = υ1 (∞) + [υ1 (0) − υ1 (∞)]e−t/τ1 = 20(1 − e−0.04t ) V

+ _

υi =

τ1 = (R1 + R2 )C = (80 + 20) × 103 × 0.25 × 10−3 = 25 s. Application of Eq. (5.96) with υ1 (0) = 0 (the capacitor had no charge prior to t = 0), υ1 (∞) = V1 = 20 V, and τ1 = 25 s leads to

i

R

0

t (s) 0

4

10

20

Figure 5-35: RC-circuit response to a 4 s long rectangular pulse.

Solution: According to Example 5-2, a rectangular pulse is equivalent to the sum of two step functions. Thus υi (t) = Vs [u(t − T1 ) − u(t − T2 )], where u(t − T1 ) accounts for the rise in level from 0 to 1 at t = T1 and the second term (with negative amplitude) serves to counteract (cancel) the first term after t = T2 . For the present problem, T1 = 0, and T2 = 4 s. Hence, the input pulse can be written as υi (t) = Vs u(t) − Vs u(t − 4).

“book” — 2015/5/4 — 7:14 — page 287 — #40

5-5

RESPONSE OF THE RL CIRCUIT

287

Since the circuit is linear, we can apply the superposition theorem to determine the capacitor response υC (t). Thus, υC (t) = υ1 (t) + υ2 (t), where υ1 (t) is the response to Vs u(t) acting alone and, similarly, υ2 (t) is the response to −Vs u(t −4) also acting alone. Response to Vs u(t) alone The response υ1 (t) is given by Eq. (5.96) with υ1 (0) = 0, υ1 (∞) = Vs , and τ = RC. Hence,

Concept Question 5-21: If Vs2 < Vs1 in the circuit of

Fig. 5-30, what would you expect the direction of the current to be after the switch is moved from position 1 to 2? Analyze the process in terms of charge accumulation on the capacitor. (See ) Exercise 5-15: Determine υ1 (t) and υ2 (t) for t ≥ 0, given that in the circuit of Fig. E5.15 C1 = 6 μF, C2 = 3 μF, R = 100 k , and neither capacitor had any charge prior to t = 0.

υ1 (t) = υ1 (∞) + [υ1 (0) − υ1 (∞)]e−t/τ = Vs (1 − e

−t/τ

)

R

(for t ≥ 0).

+ 12 V _

For Vs = 10 V and τ = RC = 25 × 103 × 0.2 × 10−3 = 5 s, υ1 (t) = 10(1 − e

−0.2t

) V

C1

υ1

C2

υ2

(for t ≥ 0). Figure E5.15

Response to −Vs u(t − 4) alone The second step function has an amplitude of −Vs and is delayed in time by 4 s. Upon reversing the polarity of Vs and replacing t with (t − 4), we have υ2 (t) = −10[1 − e−0.2(t−4) ] V

t=0

(for t ≥ 4 s).

Total response The total response for t ≥ 0 therefore is given by υC (t) = υ1 (t) + υ2 (t) = 10[1 − e−0.2t ] − 10[1 − e−0.2(t−4) ] u(t − 4) V, (5.100) where we introduced the time-shifted step function u(t − 4) to assert that the second term is zero for t ≤ 4 s. The plot of υC (t) displayed in Fig. 5-35(d) shows that υC (t) builds up to a maximum of 5.5 V by the end of the pulse (at t = 4 s) and then decays exponentially back to zero thereafter. The build-up part is due to the external excitation and often is called the forced response. In contrast, during the time period after t = 4 s, υC (t) exhibits a natural decay response as the capacitor discharges its energy into the resistor. During this latter time segment, i(t) flows in a counterclockwise direction. Concept Question 5-20: What are the three quantities needed to establish υC(t) across a capacitor in an RC circuit? (See )

Answer: υ1 (t) = 4(1 − e−5t ) V, for t ≥ 0,

υ2 (t) = 8(1 − e−5t ) V, for t ≥ 0.

5-5

(See

)

Response of the RL Circuit

With series RC circuits, we developed a first-order differential equation for υC (t), the voltage across the capacitor, and then we solved it (subject to initial and final conditions) to obtain a complete expression for υC (t). By applying iC = C dυC /dt, pC = iC υC , and wC = 21 CυC2 , we were able to determine the corresponding current passing through the capacitor, the power getting transferred to it, and the net energy stored in it. We now follow an analogous procedure for the parallel RL circuit, but our analysis will focus on the current i(t) through the inductor, instead of on the voltage across it.

5-5.1

Natural Response of the RL Circuit

After having been in the closed position for a long time, the switch in the RL circuit of Fig. 5-36(a) was moved to position 2 at t = 0, thereby disconnecting the RL circuit from the current source Is . What happens to the current i flowing through the inductor after the sudden change caused by moving the switch? That is, what is the waveform of iL (t) for t ≥ 0? To answer this question, we first note that at t = 0− (just before moving the switch), the RL circuit can be represented by the circuit in Fig. 5-36(b), in which the inductor has been

“book” — 2015/5/4 — 7:14 — page 288 — #41

288

CHAPTER 5

I0

t=0 2 + _ R0

R

+ diL υL = L dt _

+ _

R

R0

iL (t) = iL (0) e−t/τ u(t),

Initial condition at t = 0−

(c)

R0

R

L

τ=

5-5.2

+ diL υL = L dt _

Circuit at t ≥ 0 (natural response)

Figure 5-36: RL circuit disconnected from a current source at t = 0.

replaced with a short circuit. This is because under steadystate conditions iL no longer changes with time, which leads to υL = L diL /dt = 0. We also know that the current will take the path of least resistance through the short circuit. A current source entering a node connected to another node via a parallel combination of a resistor R and a short circuit will flow entirely through the short circuit. Hence, iL (0− ) = Is . Moreover, since the current through an inductor cannot change instantaneously, the initial current at t = 0 (after moving the switch) has to be iL (0) = iL (0− ) = Is . For the time period t ≥ 0, the loop equation for the RL circuit in Fig. 5-36(c) is given by diL = 0, RiL + L dt

1 L = . a R

(5.101)

(5.104)

General Form of the Step Response of the RL Circuit

To generalize our solution to the case where the RL circuit may contain sources both before and after the sudden change in the circuit configuration, we adopt the basic circuit shown in Fig. 5-37(a) in which two switches are moved simultaneously at t = 0 so as to switch the RL circuit from current source Is1 to current source Is2 . The initial state of the circuit at t = 0− (Fig. 5-37(b)) leads to the conclusion that iL (0) = iL (0− ) = Is1 . The circuit in Fig. 5-37(c) represents the arrangement at t ≥ 0. Application of KCL at the common node gives −Is2 + iR + iL = 0. Since υ is common to R and L, iR = υ/R, and by applying υL = L diL /dt, the KCL equation becomes diL + aiL = b, dt

(5.105)

where a is as given previously by Eq. (5.102) and b = aIs2 =

which can be cast in the form diL + aiL = 0, dt

(5.103)

where for the RL circuit, the time constant is given by

iL

I0

(5.102)

(natural response discharging)

+ L υL(0−) = 0 _

2

R . L

The form of Eq. (5.101) is identical to that of Eq. (5.70) for the source-free RC circuit, except that now the variable is iL (t), whereas then it was υL (t). By analogy with the solution given by Eq. (5.78), our solution for iL (t) is given by

iL(0−) = Is

1

(b)

L

a=

Switch is moved at t = 0

(a)

I0

where a is a temporary constant given by

iL

1

RC AND RL FIRST-ORDER CIRCUITS

R Is . L 2

(5.106)

Not surprisingly, Eq. (5.105) has the same form as Eq. (5.88) for the RC circuit and therefore exhibits a solution analogous

“book” — 2015/5/4 — 7:14 — page 289 — #42

5-5

RESPONSE OF THE RL CIRCUIT

289 If the sudden change in the circuit configuration happens at t = T0 instead of at t = 0, the general expression for iL (t) becomes

i

S1 1 t=0 2 Is 1

S2 R0 Is 2

2 t=0 1

+

R

L

R0

di υL= L L dt _

iL(0−) = Is1

S1 1

Is 1

S2 R0 Is 2

+

υL(0−)=0

R

_

R0

(b) Initial condition at t = 0− iL

S1 1 2 Is 1

S2 R0

2 iR 1

Is 2

(5.108)

where iL (T0 ) is the current at T0 . This expression is the analogue of Eq. (5.98) for the voltage across the capacitor.

Parallel RL Circuit Solution

2 1

· u(t − T0 ), (switch action at t = T0 )

(a)

2

  iL (t) = iL (∞) + [iL (T0 ) − iL (∞)]e−(t−T0 )/τ

R

R0

2: Analyze the circuit to determine iL (∞), the current through the inductor long after the switch action.

+ L

1: If switch action is at t = 0, analyze circuit at t = 0− (by replacing L with a short circuit) to determine initial conditions iL (0− ) and υL (0− ). Use this information to determine iL (0) and iL (0), at t immediately after the switch action. Remember that the current through an inductor cannot change instantaneously (between t = 0− and t = 0), but the voltage can.

υL

_

(c) At t ≥ 0 (natural response) Figure 5-37: RL circuit switched between two current sources at t = 0.

to the expression given by Eq. (5.96). Thus, the general form for the current through an inductor in an RL circuit is given by

iL (t) = iL (∞) + [iL (0) − iL (∞)]e−t/τ u(t), (5.107)

3: Determine the time constant τ = L/R. 4: Incorporate the information obtained in the previous three steps in Eq. (5.107):

iL (t) = iL (∞) + [iL (0) − iL (∞)]e−t/τ u(t). 5: If the switch action is at t = T0 instead of t = 0, replace 0 with T0 everywhere and use Eq. (5.108):   iL (t) = iL (∞) + [iL (T0 ) − iL (∞)]e−(t−T0 )/τ u(t−T0 ).

Example 5-13: Circuit with Two RL Branches

(switch action at t = 0)

After having been in position 1 for a long time, the SPDT switch in Fig. 5-38(a) was moved to position 2 at t = 0. Determine i1 , i2 , and i3 for t ≥ 0, given that Vs = 9.6 V, Rs = 4 k , R1 = 6 k , R2 = 12 k , L1 = 1.2 H, and L2 = 0.36 H.

with time constant τ = L/R. For the specific circuit in Fig. 5-37(a), iL (0) = Is1 and iL (∞) = Is2 .

Solution: We start by examining the initial state of the circuit before moving the switch. At t = 0− , the inductors behave

“book” — 2015/5/4 — 7:14 — page 290 — #43

290

CHAPTER 5

i1 R1

Rs

L1

i1(0 −)

i2

+ _

Vs 1

R1

R2

i3

i2(0 −)

V

R2

Rs + _

Vs

L2 L1

2

RC AND RL FIRST-ORDER CIRCUITS

L2

1

t=0

(a)

Circuit with 2 inductors

Initial condition at t = 0−

(b) i(t)

1.2 mA i1 R1

i2

1.0 mA R2

i3

i3(t)

0.8 mA 0.6 mA

L1

i2(t)

L2 0.4 mA 2

0.2 mA 0

Circuit after t = 0

(c)

i1(t) 0

0.1

0.2

0.3

0.4

t (ms)

Currents i1, i2, and i3

(d) Figure 5-38: Circuit for Example 5-13.

like short circuits, resulting in the equivalent circuit shown in Fig. 5-38(b). Application of KCL to node V gives V − Vs V V + + = 0, R1 Rs R2 whose solution is R 1 R2 Vs R1 R2 + R 1 Rs + R 2 R s 6 × 12 × 9.6 = = 4.8 V. 6 × 12 + 6 × 4 + 12 × 4

V =

Hence, the initial currents i1 (0) and i2 (0) are given by i1 (0) = i1 (0− ) =

V 4.8 = = 0.8 mA R1 6 × 103

i2 (0) = i2 (0− ) =

V 4.8 = = 0.4 mA. R2 12 × 103

and

The circuit in Fig. 5-38(c) represents the natural response circuit condition after t = 0. Even though we have two resistors and two inductors in the overall circuit, it can be treated

“book” — 2015/5/4 — 7:14 — page 291 — #44

5-5

RESPONSE OF THE RL CIRCUIT

291

as two independent RL circuits because each RL branch is connected across a short circuit. In both cases, the inductors will dissipate their magnetic energy (that they had stored prior to moving the switch) through their respective resistors. Hence, i1 (∞) = i2 (∞) = 0. The complete expressions for i1 (t) and i2 (t) for t ≥ 0 then are given by

R iL + _

υs

i1 (t) = [i1 (∞) + [i1 (0) − i1 (∞)]e−t/τ1 ] = 0.8e−t/τ1 u(t) mA

(a) RL circuit

and

υs

i2 (t) = [i2 (∞) + [i2 (0) − i2 (∞)]e−t/τ2 ] = 0.4e−t/τ2 u(t) mA,

L1 1.2 = = 2 × 10−4 s R1 6 × 103

τ2 =

L2 0.36 = = 3 × 10−5 s. R2 12 × 103

4r(t)

12 V

where τ1 and τ2 are the time constants of the two RL circuits, namely τ1 =

L

0 1

2

3

t (ms)

4

and

−12 V

(b) υs(t) = 4r(t) − 4r(t) u(t − 3 ms)

The current flowing through the short circuit is simply i3 = i1 + i2 = (0.8e−t/τ1 + 0.4e−t/τ2 ) u(t) mA.

−4r(t) u(t − 3 ms)

iL ( μA) 100 i1(t)

Example 5-14: Response to a Triangle Excitation

iL(t) = i1 + i

2

The source voltage in the circuit of Fig. 5-39(a) generates a triangular ramp function that starts at t = 0, rises linearly to 12 V at t = 3 ms, and then drops abruptly down to zero. Additionally, R = 250 , / L = 0.5 H, and no current was flowing through L prior to t = 0. (a) Synthesize υs (t) in terms of unit step functions and plot it.

0

1

2

3

4

5

6

t (ms)

i2(t)

−100 (c) iL(t) = i1(t) + i2(t)

(b) Develop the differential equation for iL (t) for t ≥ 0. Figure 5-39: Circuit and associated plot for Example 5-14.

(c) solve the equation and plot iL (t) for t ≥ 0. Solution: (a) The waveform of υs (t) shown in Fig. 5-39(b) can be synthesized as the sum of two ramp functions: υs (t) = 4r(t) − 4r(t) u(t − 3 ms)

−υs + RiL + L

diL = 0, dt

which can be rearranged into the form

= 4t u(t) − 4t u(t) u(t − 3 ms) = 4t u(t) − 4t u(t − 3 ms) V.

(b) For t ≥ 0, the KVL loop equation is given by

(5.109)

diL υs + aiL = , dt L

(5.110)

“book” — 2015/5/4 — 7:14 — page 292 — #45

292

CHAPTER 5

where a = R/L. Since υs (t) is composed of two components, we will write iL (t) as the sum of two components, iL (t) = i1 (t) + i2 (t),

(5.111)

where i1 (t) is the solution of Eq. (5.110) with υs = 4t u(t) acting alone and i2 (t) is the solution of Eq. (5.110) with υs = −4t u(t − 3 ms) acting alone. That is, 4t di1 + ai1 = = bt dt L

for t ≥ 0

(5.112a)

for t ≥ 3 ms

(5.112b)

and −4t di2 + ai2 = = −bt dt L



e

di1  + ai1 eat dt 

0

(1) The forcing function for i1 (t) is bt whereas the forcing function for i2 (t) is −bt. (2) The temporal domain of applicability for i2 (t) starts at t = 3 ms, instead of at t = 0. Hence, Eq. (5.116) can be adapted to i2 by replacing b with −b and changing the lower limit of integration to 3 ms, which gives   t i2 eat 



t



dt =

 at 

bt e



dt .

(5.113)

0

=−

eat

0



di1  + ai1 eat dt 



dt  =

t  0



d  (i1 eat ) dt  dt 

  t = i1 eat  , 0

(5.114)

and for the right-hand side, t

 at 

bt e 0

t  b at   dt = 2 e (at − 1) . a 0 

(5.115)

In view of Eqs. (5.114) and (5.115), Eq. (5.113) becomes t t  b at   at   (5.116) i1 e  = 2 e (at − 1) , 0 a 0 which leads to b at [e (at − 1) + 1]. i1 (t) e − i1 (0) = a2 at

(5.119)

b [(at − 1) + e−at ] a2

(for t ≥ 0).

(5.120)

When we apply superposition, we apply the same initial condition to both RL circuits (corresponding to the two components of υs (t)). Thus, i1 (0) = i2 (3 ms) = 0, and Eq. (5.120) simplifies to i2 (t) = −

b [(at − 1) − (0.003a − 1)e−a(t−0.003) ] a2 (for t ≥ 3 ms). (5.121)

Total solution for iL (t) For R = 250 and L = 0.5 H, a = R/L = 500, b = 4/L = 8, and  for 0 ≤ t ≤ 3 ms, i1 (t) iL (t) = i1 (t) + i2 (t) for t ≥ 3 ms, ⎧ 32[(500t − 1) + e−500t ] μA ⎪ ⎪ ⎪ ⎨ for 0 ≤ t < 3 ms, = (5.122) −500t μA ⎪ 103.7e ⎪ ⎪ ⎩ for t ≥ 3 ms. Figure 5-39(c) displays a plot of iL (t) versus t.

(5.117) Concept Question 5-22: Compare Eq. (5.96) with

given that i1 (0) = 0, the expression for i1 (t) becomes i1 (t) =

t −b at    e (at − 1) ,  3 ms a2

b at [e (at − 1) − e0.003a (0.003a − 1)]. a2

For the left-hand side, t 

=

i2 (t) eat − i2 (3 ms) e0.003a

We start by multiplying both sides of Eq. (5.112a) by eat and then integrating from 0 to t: at 

Equations (5.112a) and (5.112b) are identical in form, except for two important differences:

which leads to

Current i1 (t) alone

t 

Current i2 (t) alone

3 ms

with b = 4/L.

RC AND RL FIRST-ORDER CIRCUITS

(5.118)

Eq. (5.107) to draw an analogy between RC and RL circuits. υC, R, and C of the RC circuit correspond to which parameters of the RL circuit? (See )

“book” — 2015/5/4 — 7:14 — page 293 — #46

TECHNOLOGY BRIEF 13: HARD DISK DRIVES (HDD)

Technology Brief 13 Hard Disk Drives (HDD) Although invented in 1956, the hard disk drive (HDD) arguably is still the most commonly used data-storage device among nonvolatile storage media available today. It is the availability of vast amounts of relatively inexpensive hard-drive space that has made search engines, webmail, and online games possible. Over the past 40 years, improvements in HDD technology have led to huge increases in storage density, which are simultaneous with the significant reduction in physical size. The term hard disk or hard drive evolved from common usage as a means to distinguish these devices from flexible (floppy) disk drives.

HDD Operation Hard drives make use of magnetic material to read and write data. A nonmagnetic disc ranging in diameter from 36 to 146 mm is coated with a thin film of magnetic material, such as an iron or cobalt alloy. When a strong magnetic field is applied across a small area of the disc, it causes the atoms in that area to align along the orientation of the field, providing the mechanism for writing bits of data onto the disc (Fig. TF13-1). Conversely, by detecting the aligned field, data can be read back from the disc. The hard drive is equipped with an arm that can be moved across the surface of the disc (Fig. TF13-2), and the disc itself is spun around to make all of the magnetic surface accessible to the writing or reading heads. The reading

293 and writing elements are physically moved along the radius of the disk by using a magnet with a coil wrapped around it. When current is driven into the coil, it produces a magnetic force that moves the actuator. Because writing onto or reading from the magnetized surface can be performed very rapidly (fraction of a microsecond), hard drives are spun at very high speeds (5,000 to 15,000 rpm) when directed to record or retrieve information. Amazingly, hard-drive heads usually hover at a height of about 25 nm above the surface of the magnetic disc while the disc is spinning at such high speeds! The extremely small gap between the head and the disc is maintained by having the head “ride” on a thin cushion of air trapped between the head and the surface of the spinning disc. To prevent accidental scratches, the disc is coated with carbon- or Teflon-like materials. Hard drives are packaged carefully to prevent dust and other airborne particles from interfering with the drive’s operation. In combination with the air motion caused by the spinning disc, a very fine air filter is used to keep dust out while maintaining the air pressure necessary to cushion the spinning discs. Hard drives intended for operation at high altitudes (or low air pressure) are sealed hermetically so as to make them airtight.

Modern Drive Technology Early hard drives performed read and write operations by using an inductor coil placed at the tip of the head. When electric current is made to flow through the coil, the coil induces a magnetic field which in turn aligns the

Standard Magnetic Recording

Spindle Actuator arm

Perpendicular Magnetic Recording Heads

Figure TF13-2: Close-up of a disassembled hard drive Figure TF13-1: Longitudinal and perpendicular writing techniques.

showing the magnetic discs mounted on a spindle and an actuator arm. The head sits at the end of the arm and performs the read/write operations as the disc spins.

“book” — 2015/5/4 — 7:14 — page 294 — #47

Recent Developments A new wave of developments is pushing hard drives into the tens of terabytes. Already in commercial use is shingled magnetic recording (SMR). Conventional drives write bits in parallel rows Fig. TF13-3(a)), usually with a slight gap between them. Making the individual track width smaller is extremely difficult because, as mentioned above, very small magentic grains are not stable (or, conversely, to make very small grains stable makes them very hard to read/write with a magnetic head). The SMR solution (Fig. TF13-3(b)) is to lay bits down in overlapping tracks, exactly like roof shingles (where each shingle row

Track n Track n + 1 Track n + 2

Track pitch

Cross track

Down track (direction of rotation)

Writer and reader gap widths

(a) Schematic of conventional magnetic recording Track n Track n + 1 Track n + 2

Band A

atoms of the magnetic material (i.e., a write operation). The same coil also is used to detect the presence of aligned atoms, thereby providing the read operation. The many major developments that shaped the evolution of read/write heads over the past 50 years have introduced two major differences between the modern hard-drive heads and the original models. Instead of using the same head for both reading and writing, separate heads are now used for the two operations. Furthermore, the writing operation is now carried out with a lithographically defined thin-film head, thereby reducing the feature size of the head by several orders of magnitude. The feature size is the area occupied by a single bit on the disc surface, which is determined in part by the size of the write head. Decreasing feature size leads to increased recording density. The read operation—housed separately next to the write head—uses a magnetoresistive material whose resistance changes when exposed to a magnetic field—even when the field intensity is exceedingly small. In modern hard drives, high magnetoresistive sensitivities are realized through the application of either the giant magnetoresistance (GMR) phenomena or the tunneling magnetoresistance (TMR) effect exhibited by certain materials. The 2007 Nobel prize in physics was awarded to Albert Fert and Peter Grunberg ¨ for their discovery of GMR. A consequence of the extremely small size of the magnetic bits (each bit in a 100-Gb/in2 disc is about 40 nm long) is that temperature variations can lead to loss of information over time. One method developed to combat this issue is to use two magnetic layers separated by a thin (∼ 1 nm) insulator, which increases the stability of the stored bit. Another recent innovation that is already in production involves the use of perpendicular magnetic recording (PMR) as illustrated in Fig. TF13-1. PMR makes it possible to align bits more compactly next to each other.

TECHNOLOGY BRIEF 13: HARD DISK DRIVES (HDD)

Track n + 3 Track n + 4 Track n + 5

Band B

294

(b) Schematic of shingled magnetic recording FigureTF13-3: Schematics of (a) conventional magnetic recording and (b) shingled magnetic recording with two 3-track bands.

sits slightly on top of one adjacent row and slightly below the other). The advantage is that the size of the track (and hence, the grain), does not change but the overall density increases. This works because a magnetic head can still read the state of the magentic grain even if it slightly overlapped with a nearby grain. The difficulty of this method is that the writing process slows down since every time we write to one of the overlapped rows, we must also rewrite the neighboring rows. The tracks are organized into bands (Fig. TF13-3(b)) and each band is thus rewritten as needed. Coordinating this write activity can be handled in firmware on the drive itself or in the computer’s operating system (if it has the appropriate driver to handle such drives). A variety of other techniques (including the GMR heads discussed above) are being explored to increase areal density; in general, these focus on allowing smaller grains by making them harder to write magnetically (which makes them consequently more temperature stable). Among these are heat-assisted, microwave-assisted and patterning single-grain (or close to single-grain) isolated magnetic islands (instead of a continuous magnetic thin film); this is known as bit-patterned media (BPM). It is estimated that techniques such as these will enable densities on the order of 1–10 Tb/in2 in the next decade.

“book” — 2015/5/4 — 7:14 — page 295 — #48

5-6

RC OP-AMP CIRCUITS

295

Concept Question 5-23: Suppose the switch in the circuit of Fig. 5-36(a) had been open for a long time, and then it was closed suddenly. Will Is initially flow through R or L? (See )

RC Integrator υC = υout i C _

+

iR

_

υn

Exercise 5-16: Determine i1 (t) and i2 (t) for t ≥ 0

given that, in the circuit of Fig. E5.16, L1 = 6 mH, L2 = 12 mH, and R = 2 . Assume −

in = 0

R

υi

+ _

υp

C υout

+

RL



i1 (0 ) = i2 (0 ) = 0. Figure 5-40: Integrator circuit.

i1

t=0 1.8 A

R

i2

L1

L2

Figure E5.16

The ideal op-amp model has two constraints. The voltage constraint states that υp = υn , and since υp = 0 in the circuit of Fig. 5-40, it follows that υn = 0. Hence, the current iR flowing through R is given by iR =

Answer: i1(t) = 1.2(1 − e−500t ) u(t) A,

i2(t) = 0.6(1 − e−500t ) u(t) A. (See

C3

)

5-6.1

Ideal Op-Amp Integrator

The circuit shown in Fig. 5-40 resembles the standard invertingamplifier circuit of Section 4-4, except that its feedback resistor Rf has been replaced with a capacitor C, converting it into an op-amp integrator. As we show shortly:  The output voltage υout of the RC integrator circuit is directly proportional to the time integral of the input signal υi . 

(5.123)

Given that υn = 0, the voltage υC across C is simply υout , and the current flowing through it is

5-6 RC Op-Amp Circuits Adding capacitors and inductors to resistive circuits vastly expands their utility and versatility. In this section, we consider a few examples of circuits in which capacitors are used in conjunction with op amps to perform integration, differentiation, and related operations. Even though these specific functions also can be realized through the use of inductors, capacitors are usually the preferred option (whenever such a choice is possible) because of their smaller physical size and availability in planar form.

υi . R

iC = C

dυout . dt

(5.124)

At node υn , iR + iC − in = 0.

(5.125)

In view of the second op-amp constraint, namely in = ip = 0, it follows that iC = −iR

(5.126)

or

1 dυout (5.127) =− υi . dt RC Upon integrating both sides of Eq. (5.127) from an initial reference time t0 to time t, we have t  t0

dυout dt 



1 dt = − RC 

t

υi dt  ,

(5.128)

t0

which leads to 1 υout (t) = − RC

t t0

υi (t  ) dt  + υout (t0 ).

(5.129)

“book” — 2015/5/4 — 7:14 — page 296 — #49

296

CHAPTER 5

Time t0 is the time at which the integration process begins, and υout (t0 ) is the initial voltage across the capacitor at that instant in time. Thus, according to Eq. (5.129), the output voltage (which is also the voltage across the capacitor) is equal to whatever voltage existed across the capacitor at the start of the integration process, υout (t0 ), incremented by an amount equal to the integrated value of the input voltage (from t0 to present time t) and multiplied by a (negative) scaling factor (−1/RC).

RC AND RL FIRST-ORDER CIRCUITS

υi (V) Input 3 0

t (s) 1 2

3 4

5

6

−3 (a) υout (V)

 Since the magnitude of the output voltage, |υout |, cannot exceed the supply voltage Vcc , the values of R and C have to be chosen carefully so as to avoid saturating the op amp. 

12

Output when Vcc = 14 V

6

If the time scale can be conveniently chosen such that the initial reference time t0 = 0 and the capacitor was uncharged at that point in time (i.e., υout (0) = 0), then Eq. (5.129) simplifies to

0

t (s) 1

2

3

4

5

6

−6 −12 (b)

1 υout (t) = − RC

t

υi (t  ) dt  (if υout (0) = 0).

υout (V)

(5.130)

0

12

Output when Vcc = 9 V

6 0

Example 5-15: Square-Wave Input Signal

The square-wave signal shown in Fig. 5-41(a) is applied at the input of an ideal integrator circuit with an initial capacitor voltage of zero at t = 0. If R = 200 k and C = 2.5 μF, determine the waveform of the corresponding output voltage for an amp with (a) Vcc = 14 V and (b) Vcc = 9 V. Solution: (a) The scaling factor is given by −

−6 −9 −12

t (s) 1

2

3

4

5

6

Clipped output

(c) Figure 5-41: Example 5-15 (a) input signal, (b) output signal with no op-amp saturation, and (c) output signal with op-amp saturation at −9 V.

1 1 = −2 s−1 . =− 5 RC 2 × 10 × 2.5 × 10−6

For the time period 0 ≤ t ≤ 2 s (first half of the first cycle), t υout (t) = −2 0

υi dt  = −2

t

3 dt  = −6t V

0

(0 ≤ t ≤ 2 s), which is represented by the first ramp function shown in Fig. 5-41(b). The polarity reversal of υi during the second half

of the first cycle causes the energy that had been stored in the capacitor to be discharged, concluding the cycle with no net voltage across the capacitor. The process then is repeated during succeeding cycles. We note that because |υout | never exceeds |Vcc | = 14 V, no saturation occurs in the op amp. (b) For the op amp with Vcc = 9 V, the waveform shown in Fig. 5-41(c) is the same as that in Fig. 5-41(b), except that it is clipped at −9 V.

“book” — 2015/5/4 — 7:14 — page 297 — #50

5-6

RC OP-AMP CIRCUITS

297

Example 5-16: Pulse Response of an Op-Amp Circuit

RC Differentiator R iC

C

in = 0 υn

υi

+ _

υp

iR

_

υout

+

Solution: One possible approach to solving the problem is to analyze the circuit twice—once for the duration of the pulse (0 to 0.3 s) and a second time for t > 0.3 s. An alternative approach is to synthesize the rectangular pulse as the sum of two step functions, to seek an independent solution for each step function, and then to add up the solutions (superposition). We will illustrate both methods.

RL

Figure 5-42: Differentiator circuit.

5-6.2

(a) Method 1: Two Time Segments

Ideal Op Amp Differentiator

The integrator circuit of Fig. 5-40 can be converted into the differentiator circuit of Fig. 5-42 by simply interchanging the locations of R and C. For the differentiator circuit, application of the voltage and current constraints leads to dυi , iC = C dt

υout iR = , R

The op-amp circuit shown in Fig. 5-43(a) is subjected to an input pulse of amplitude Vs = 2.4 V and duration T0 = 0.3 s. Determine and plot the output voltage υout (t) for t ≥ 0, assuming that the capacitor was uncharged before t = 0.

Time Segment 1: At node υn ,

i1 + i2 + i3 = 0, or, using the node voltage method,

and

iC = −iR .

d υn − υout1 υn − Vs +C = 0, (υn − υout1 ) + R1 dt R2

Consequently, υout

dυi = −RC , dt

(5.131)

where υout1 is the output voltage during time segment 1. Since υp = 0, injection of the ideal op-amp voltage constraint υp = υn leads to

which states that:  The output voltage of the differentiator circuit is proportional directly to the time derivative of its input voltage υi , and the proportionality factor is (−RC). The differentiator circuit performs the inverse function of that performed by the integrator circuit. 

5-6.3

0 ≤ t ≤ 0.3 s, and υi = Vs = 2.4 V.

Other Op-Amp Circuits

The relative ease with which we were able to develop input– output relationships for the ideal integrator and differentiator circuits is attributed (at least in part) to the relative simplicity of those circuits. Aside from the load resistor RL (which exercised no influence on the solutions), the circuits in Figs. 5-40 and 5-42 consisted each of one resistor and one capacitor. Now, through two examples, we demonstrate ways to approach the analysis of RC op-amp circuits that may have more complicated architectures.

C

υout1 Vs dυout1 + =− , dt R2 R1

which can be cast in the standard first-order differentialequation form given by dυout1 + aυout1 = b, dt

(5.132)

where a=

1 , R2 C

and

b=−

Vs . R1 C

Equation (5.132) is analogous to Eq. (5.88), so its solution is analogous to that given by Eq. (5.94), namely b (1 − e−at ) a V s R2 − (1 − e−t/τ ), R1

υout1 (t) = υout1 (0) e−at + = υout1 (0) e−t/τ

(5.133)

“book” — 2015/5/4 — 7:14 — page 298 — #51

298

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

R2 = 10 kΩ

in = 0

R1 = 2 kΩ Vs

i1 υn υp

+ _

υi(t) = t=0

C = 25 μF

i2

i3

+υC_ _ υout

+

t = 0.3 s (a) Op-amp circuit

υout (V) 0

0

0.3

0.5

1

1.5

2

t (s)

−2 Capacitor discharging

−4 −6 −8

Capacitor building up charge −10 (b) υout(t) Figure 5-43: Op-amp circuit of Example 5-16.

where 1 τ = = R2 C = 0.25 s. a Given that υn = 0, it is evident from the circuit in Fig. 5-43(a) that υout1 = −υC1 , where υC1 is the voltage across the capacitor during the first time segment. According to the problem statement, the initial condition υC1 (0− ) = 0, and since the voltage across a capacitor cannot change instantaneously, it follows that υout1 (0) = −υC1 (0) = −υC1 (0− ) = 0.

Upon incorporating this piece of information into our solution, we have the natural response υout1 (t) = −

Vs R2 (1 − e−t/τ ) R1

= −12(1 − e−4t ) V Time Segment 2:

(for 0 ≤ t ≤ 0.3 s). (5.134)

t > 0.3 s, and υi = 0.

The form of the solution for this time segment is the same as that given by Eq. (5.133) for the preceding time segment, except for three modifications: (a) The input voltage is now zero, so we should set Vs = 0.

“book” — 2015/5/4 — 7:14 — page 299 — #52

5-6

RC OP-AMP CIRCUITS

299

(b) The time variable t should be replaced with (t − 0.3 s) to reflect the fact that our starting (reference) time is t = 0.3 s, not t = 0. (c) The initial voltage υout2 (0.3 s) is not zero (because the capacitor had been building up charge during the previous time segment). Hence, for time segment 2, υout2 is given by υout2 (t) = υout2 (0.3) e−4(t−0.3)

(for t > 0.3 s).

The initial voltage υout2 (0.3) is equal to the voltage that existed during the previous time segment at t = 0.3 s. Hence, υout2 (0.3) = υout1 (0.3) = −12(1 − e

−4×0.3

) = −8.4 V.

Hence, υout2 (t) = −8.4e

−4(t−0.3 s)

V

(for t > 0.3 s).

(5.135)

The combined output response to the input pulse is displayed in Fig. 5-43(b). (b) Method 2: Two Step Functions

In view of the definition of the step function, the complete solution is given by υout (t) = υouta (t) + υoutb (t)  υouta (t) for 0 ≤ t ≤ 0.3 s = υouta (t) + υoutb (t) for t > 0.3 s.

It is a relatively straightforward exercise to demonstrate that the two methods do indeed provide the same solution. Example 5-17: Op-Amp Circuit with Output Capacitor

Determine υC (t), the voltage across the capacitor in Fig. 5-44(a), given that υi (t) = 3u(t) V, the capacitor had no charge on it prior to t = 0, R1 = 1 k , R2 = 15 k , R3 = 30 k , R4 = 12 k , R5 = 24 k , and C = 50 μF. Solution: The capacitor is on the output (load) side of the op amp, so one possible approach to solving the problem is to (a) temporarily replace the capacitor with an open circuit; (b) determine the Th´evenin equivalent circuit at terminals (a, b); and (c) reinsert the capacitor as in Fig. 5-44(c) and analyze the circuit.

By modeling the rectangular pulse as υi (t) = Vs [u(t) − u(t − 0.3 s)],

(5.136)

To that end, we start by relating υout to υi . Given that for the ideal op amp υn = υp and ip = 0, it follows that

we can develop a generic solution to a step-function input and then use it to find υout (t) = υouta (t) + υoutb (t). We will treat the two step functions as two independent sources, and we will apply the same initial-condition information to both cases; that is, when treating the case of the second step function, we do so as if the first step function had never existed. To that end, the response of the first step function is given by Eq. (5.134) as υouta (t) = −12(1 − e−4t ) u(t) V

(5.139)

(for t ≥ 0). (5.137)

Similarly, after reversing the polarity of Vs and incorporating a time delay of 0.3 s, υoutb (t) = 12(1−e−4(t−0.3) ) u(t −0.3) V

(for t ≥ 0.3 s). (5.138)

υn = υp = υi . Moreover, since in = 0, υn and υout are related by a voltage divider between nodes c and d:     R2 + R3 R2 + R3 υn = υi . υout = R2 R2 With the capacitor removed, the Th´evenin voltage across terminals (a, b) in Fig. 5-44(a) is equal to the voltage across R5 , which is related to υout by the voltage-division rule  υTh =

R5 R4 + R 5

 υout

  R2 + R3 R5 υi R4 + R 5 R2    24 15 + 30 = × 3 = 6u(t) V 12 + 24 15 

=

(for t ≥ 0).

“book” — 2015/5/4 — 7:14 — page 300 — #53

300

CHAPTER 5

R1

(a) Op-amp circuit υi = 3 u(t)

υp ip = 0 υn in = 0

+ _

+

RC AND RL FIRST-ORDER CIRCUITS

R4

c υout

_

a

R3 R5 R2

υp = υn

R4

c

a

R3 Ro

RTh

R5 R2 b

d RTh (c) Equivalent circuit

+ υ _C

b

d

(b) Relevant circuit for finding RTh , with op amp replaced with its output resistance Ro

C

υTh = 6u(t)

a

+ _

C b

+ υ _ C

Figure 5-44: Circuit for Example 5-17.

Our next task is to determine the value of RTh . To that end, we set υi = 0. Consequently, υp − υn = 0, in which case the op-amp’s equivalent circuit at terminals (c, d) consists of only its output resistance R0 . Figure 5-44(b) contains the relevant part of the overall circuit seen by terminals (a, b). For the real op amp, R0 is on the order of 10 to 100 , which is at least two orders of magnitude smaller than any of the other resistors in the circuit, lending justification to the ideal op-amp model which sets R0 = 0 (thereby shorting out (R2 + R3 )). Consequently, RTh

R 4 R5 12 × 24 = R4  R5 = = = 8 k . R4 + R 5 12 + 24

With υTh and RTh known, we now have a circuit (Fig. 5-44(c)) that resembles the step-function circuit of Fig. 5-30(a). Its solution is given by Eq. (5.97) using Vs1 = 0 and Vs2 = Vs , namely υC (t) = Vs (1 − e−t/τ ).

In the present case, Vs = υTh = 6 V, and τ = RTh C = 8 × 103 × 50 × 10−6 = 0.4 s. The capacitor response is therefore given by υC (t) = 6(1 − e−2.5t ) u(t) V.

Example 5-18: Differential Equation Solver

Design an op-amp circuit whose output is the solution of the differential equation d 2υ dυ +8 + 2υ = 4υs (t), dt 2 dt where υs (t) is a sinusoidal source given by υs (t) = 3 sin(200t) u(t).

(5.140)

“book” — 2015/5/4 — 7:14 — page 301 — #54

TECHNOLOGY BRIEF 14: CAPACITIVE SENSORS

Technology Brief 14 Capacitive Sensors Capacitive sensors are used to convert information from the real world to a change in capacitance that can be detected by an electric circuit. Even though capacitors can assume many different shapes, the basic concepts can be easily explained using the shape and properties of the parallel plate capacitor, for which the capacitance C is given by εA C= , d where ε is the permittivity of the material between the plates, A is the area of each plate, and d is the spacing between the plates. So, most capacitive sensors operate by measuring the change in one or more of these three basic parameters, in response to external physical stimuli. Let us examine each one of these three parameters separately and how it can be used to measure external stimuli.

Applications Based on Change in Permittivity ε The electrical permittivity ε of a given material is an inherent property of that material; its value is dictated

Table TT14-1: Relative permittivity εr of common materials.a

ε = εr ε0 and ε0 = 8.854 × 10−12 F/m Material Vacuum Air (at sea level) Low Permittivity Materials Styrofoam Teflon Petroleum oil Wood (dry) Paraffin Polyethylene Polystyrene Paper Rubber Plexiglass Glass Quartz Water Biological Materials a These

Relative Permittivity, εr 1 1.0006 1.03 2.1 2.1 1.5–4 2.2 2.25 2.6 2–4 2.2–4.1 3.4 4.5–10 3.8–5 72–80 40–70

are at room temperature (20 ◦ C).

301 by the polarization behavior of that material’s molecular structure, relative to the absence of polarizability (as in free space or vacuum). In free space, ε = ε0 = 8.854 × 10−12 F/m, and for all other media, it is convenient to express the permittivity of a material relative to that for free space through the relative permittivity εr = ε/ε0 . Table TT14-1 provides a list for various types of materials. We note that for plastic, glass, and most ceramics, εr is in the range between 2 and 4, which makes them different (electrically) from air (εr = 1 for air), but not markedly so. In contrast, water-based materials—such as biological materials or parts of the body—have an εr in the range of 60–80, making them electrically very different from both air and dry materials. This means that their presence can be easily detected by a capacitive sensor, which is the basis of capacitive touchscreens, fluid and moisture meters, and some proximity meters. Capacitive Touch Buttons An example of a capacitive touch sensor is shown in Fig. TF14-1. The capacitor has two conducting surfaces labeled sensor pad and ground hatch. In general, the two conductors are separated either vertically or horizontally, and covered with a layer of glass or plastic. By applying a voltage source (supplied by the printed circuit board) between the conducting surfaces, electric field lines get established between them. When no finger (or a capacitive stylus) is present near the sensor pad, the electric field lines flow through the glass or plastic cover, but when in the proximity of a finger, the electric field lines pass partially through the finger, and since the finger has a relative permittivity comparable to that of water, its

Overlay Ground hatch

Sensor pad Ground hatch

PC board

Figure TF14-1: A capacitive touch sensor uses the high permittivity of the finger to change the capacitance. The finger does not need to come in direct contact with the sensor in order to be detected.

“book” — 2015/5/4 — 7:14 — page 302 — #55

302

TECHNOLOGY BRIEF 14: CAPACITIVE SENSORS

Contact pad

Sensing film

Alumina substrate Interdigitated electrode Figure TF14-2: Interdigitated humidity sensor. (Credit: Hygrometrix.)

proximity changes the overall capacitance of the circuit. The electric field starts on one of the conductors and ends on the other, basically making an arc between them. When the finger comes near either one or both of the two conductors, it changes this field (note the electric field arrow pointing straight up at the finger, which would not be there without the finger), and this in turn changes the capacitance. Another way to think about the process is in terms of the electric charge stored at the two conductors. The presence of the finger changes the effective permittivity of the medium through which the electric field lines flow, thereby changing the effective capacitance C. Since for any capacitor, C = Q/V —where Q is the charge on the conductor connected to the positive terminal of the voltage source and V is the voltage of the source—it follows that increasing C leads to an increase in Q (with V remaining constant). Hence, when the finger approaches the sensor pad, additional charge accumulates at the two conductors (with more +Q at the sensor pad and a corresponding −Q at the ground hatch). Humidity Sensor Another example of a capacitive sensor that also relies on measuring the change in permittivity is the humidity sensor featured in Fig. TF14-2. A sensing film absorbs moisture from the air, thereby changing the capacitance of the interdigitated line in proportion to the humidity in the air surrounding the sensor. “Seeing” through Walls The capacitive sensing technique also is used to “see” inside boxes, through walls, or through basically any low-conductivity low-permittivity material (paper, plastic, glass, etc.). An example is illustrated in Fig. TF14-3, in which a capacitive sensor on an assembly line is used to determine if a metal object is placed inside a box. The

Figure TF14-3: Capacitive proximity sensors can “see” through low permittivity materials such as paper, cardboard, plastic, and glass and detect objects composed of a wide variety of materials including metals, fluids, etc. Here, a capacitive sensor detects the contents of a box. (Graphic courtesy of Balluff.) object does not have to be metal, but its permittivity has to be significantly different from that of the paper or plastic enclosure. A similar application of capacitive sensors is to locate wooden studs through plaster walls. Fluid Gauge Capacitive sensors can serve as fluid gauges by measuring the height of a fluid in a tank or reservoir. Examples include gasoline and oil level gauges used in cars. If the tank is made of plastic or glass, metal strips on the outside of the tank can determine the height of the fluid without having to make contact with the fluid. This is very useful when the fluid is caustic or sterile. If the tank is metal, the strips must be placed inside. In either case, the sensor consists of two capacitors, one (C2 in Fig. TF14-4) with metal plates separated by a reference fluid, and another (C1 ) in which the fluid level is a variable. If the permittivity of the fluid is ε and the height of the fluid in the upper container in Fig. TF14-4 is h, the ratio of the two capacitances is given by C1 = ah + b, C2 where a and b are known constants related to ε and the dimensions of the two capacitors. Hence, by measuring the two capacitances with an external circuit, the sensor provides a direct measurement of the fluid height h.

“book” — 2015/5/4 — 7:14 — page 303 — #56

TECHNOLOGY BRIEF 14: CAPACITIVE SENSORS

303

C1

x

W d

Air

L C1 Variable

ε

h

L

h0

Figure TF14-6: Capacitance is proportional to overlap area A = W (L − x), so when plates slide past each other the capacitance decreases in proportion to the shifted distance x.

ε

C2 Reference

x

C2

Applications Based on Change in Area A

Figure TF14-4: Fluid height can be measured from the outside of a plastic or glass tank using a pair of parallel plate capacitors on the outside of the tank.

C = ε0 εr

a×b d

Pressure

Transducer

Data

Figure TF14-5: Capactive transducer responding to pressure from a sound wave.

The change in the effective area common to the two conducting surfaces can also change the capacitance C. If one plate is slid past the other in Fig. TF14-6, the effective area A changes as a function of the shifted distance x. The capacitance is maximum when they are perfectly lined up, corresponding to x = 0, and changes approximately linearly as (L − x). This can be used to align two objects, or to determine any other manual displacement in either one or two directions. The MEMS capacitive vibration sensor shown in Fig.TF14-7 uses two interdigital electrodes, one static and another moveable. When mounted in a car, for example, car acceleration or deceleration causes the moveable electrode to respond accordingly, which changes the capacitance between the two electrodes, thereby providing the means to measure acceleration. Such a sensor is called an accelerometer.

Applications Based on Change in Inter-Conductor Distance d As noted earlier, the capacitance C is inversely proportional to the distance d between the two conductors. This dependence can be used to measure pressure, as illustrated by the diagram in Fig. TF14-5. We call such a sensor an electrical transducer because it converts one type of energy (mechanical) into another (electrical). The capacitor has one stationary conducting plate on the back side and a flexible conducting membrane on the side exposed to the incident pressure carried by an acoustic wave. The sound wave causes the membrane to vibrate, thereby changing the capacitance, which is measured and processed by an external circuit. This type of capacitive transducer is used in numerous industrial applications.

FigureTF14-7: Microelectromechanical system (MEMS) vibration sensor using interdigitated static and movable electrodes. (Credit: STMicroelectronics.)

“book” — 2015/5/4 — 7:14 — page 304 — #57

304

CHAPTER 5

The step function u(t) denotes that the source is connected to the circuit at t = 0. In your circuit, you may use a sinusoidal source of any amplitude and angular frequency.

combined by a weighted op-amp summer in which the gains can be adjusted to obtain the desired output υ. In Fig. 5-45, υ is the output of op amp 4, as well as the input to op amp 1, which is a differentiator with a gain factor of −RC = −1 (the values of R and C are selected such that their product is 1). The output of op amp 1 is simply −υ  . When followed by a second differentiator (op amp 2), we obtain υ  . Op amp 3 serves as an inverter with gain of −1. Finally, op amp 4 is a summing amplifier that performs the sum of all three terms in Eq. (5.141). The values of the resistors preceding the summing point at the input to op amp 4 are selected to provide the correct weights, namely (6R/12R) = 1/2 for υ  , (6R/1.5R) = 4 for υ  , and (6R/R) = 6 for the sinusoidal source. The switch serves to initiate the process at t = 0. Prior to that, υ = 0. To avoid saturation, the supply voltage Vcc of each op amp should exceed the maximum possible voltage at its output. If one were to construct the circuit and close the switch, the voltage υ(t) observed at the output of op amp 4 would be the same solution we would obtain were we to solve the differential equation analytically.

Solution: Using op amps, multiple circuit configurations can be constructed to solve the given differential equation. One such configuration is shown in Fig. 5-45. If in Eq. (5.140) we denote dυ/dt = υ  and d 2 υ/dt 2 = υ  and then solve for υ, we have 1  υ − 4υ  + 2υs (t) 2 1 = − υ  − 4υ  + 6 sin(200t) u(t). 2

RC AND RL FIRST-ORDER CIRCUITS

υ=−

(5.141)

One approach for designing this circuit is to realize that the output must be υ, and somehow within the circuit we will also need υ  and υ  . We can design a differentiator with a gain of 1 and feed in the υ (output), and then feed that into a second differentiator to get υ  . The values of υ  , υ  , and υs can be

R R C υ

6R

_ Op Amp 1

−υ′

C

_ Op Amp 2

+

+

12R υ′′

_

Summing point

Op Amp 4

+

Differentiator Gain = −1

Differentiator

Summer

R 1

R RC = 1 Gain = −1

υ = − 2 υ′′ − 4υ′ + 6 sin(200t)

_ Op Amp 3

+

1.5R

υ ≤ Vcc

υ′

Inverter Gain = −1 R sin(200t) t=0 Figure 5-45: Op-amp circuit whose output υ(t) is a solution to υ  + 8υ  + 2υ = 12 sin(200t) u(t).

υ

“book” — 2015/5/4 — 7:14 — page 305 — #58

5-7 APPLICATION NOTE: PARASITIC CAPACITANCE AND COMPUTER PROCESSOR SPEED Concept Question 5-24: What causes clipping of the

waveform at the output of an op-amp integrator circuit? Can clipping occur at the output of a differentiator circuit? (See ) Concept Question 5-25: If υs (t) is the input signal to

a two-stage op-amp circuit with the first stage being an integrator with R1 C1 = 0.01 and the second stage being a differentiator with R2 C2 = 0.01, under what circumstances will the output waveform υout (t) be the same or different from υs(t)? (See )

Exercise 5-17: The input signal to an ideal integrator circuit with RC = 2 × 10−3 s and Vcc = 15 V is given by υs(t) = 2 sin 100t V. What is υout (t)? Answer: υout (t) = 10[cos(100t) − 1] V. (See

C3

)

5-7 Application Note: Parasitic Capacitance and Computer Processor Speed As was noted in Section 4-11 and in Technology Brief 10, the primary computational element in modern computer processors is the CMOS transistor. How quickly a single logic gate is able to switch its output between logic states 0 and 1 determines how fast the entire processor can perform complex calculations. Figure 5-46(a) displays a sample of a digital sequence, perhaps at the output of a digital inverter. The individual pulses, each denoting a logic state of 0 or 1, are each of duration T . If it were possible to switch between states instantaneously, the maximum number of pulses that can be sequenced per 1 second is 1/T . We refer to this rate by several names, including the pulse repetition frequency, switching frequency, and clock speed. In the present case, we shall call it the switching frequency and assign it the symbol fs . That is, fs =

Exercise 5-18: Repeat Exercise 5-17 for a differentiator

instead of an integrator. Answer: υout (t) = −0.4 cos 100t V. (See

C

)

305

1 T

(Hz).

(5.142)

So if T = 1 ns, fs = 1/10−9 = 1 GHz, and if we can make the pulse duration narrower, we can increase fs accordingly.

Vout VDD

1 0 1

Logic state 1

(a) Pulses 0

Logic state 0

T

t Vout VDD

0

State 1

trise

T

State 0

tfall

t (b) Expanded view Figure 5-46: Pulse sequence.

“book” — 2015/5/4 — 7:14 — page 306 — #59

306

CHAPTER 5

Such a conclusion would be true if we can indeed arrange to have the logic circuit switch between states instantaneously, but it cannot. In Fig. 5-46(b), we show an expanded view of three pulses representing the sequence 101. We observe that the switching process is represented by ramp functions (rather than step functions) and it takes a finite amount of time for the voltage to change between a 0 state and a 1 state, which we shall call the rise time trise . Similarly, the fall time between states 1 and 0 is tfall . [The linear rise and fall responses are actually artifacts of certain simplifying assumptions. In general, the responses involve exponentials, in which case it is more appropriate to define trise and tfall as the durations between the 10 percent level and 90 percent level of the change in voltage.] The total time associated with a pulse is Ttotal = T + trise + tfall = T + 2trise

Wire capacitor l

π ε  ln[(d/2a) + (d/2a)2 − 1] πε ≈ if d a ln(d/a)

C=

Figure 5-47: Capacitance of a two-wire configuration where ε is the permittivity of the material separating the wires.

(if trise = tfall ),

and the associated switching frequency is 1 Ttotal

=

1 . T + 2trise

Even if T can be reduced to zero, the maximum possible switching speed (without overlap between adjacent pulses) would be fs (max) =

1 . 2trise

2a

d

5-7.1

fs =

RC AND RL FIRST-ORDER CIRCUITS

(5.143)

As we shall see shortly, the switching times (trise and tfall ) are governed in part by the capacitances in the circuit. Consequently, capacitances play a major role in determining the ultimate switching speed of a digital circuit. In fact, capacitances also govern the switching speeds of the wires— often referred to as the bus—that connect the processor to the various other devices on a computer motherboard.

 Whereas the processor speed of a modern computer is in the GHz range, the bus speed usually is slower by a factor of 3 to 10.  This is (in part) why a computer appears to slow down when the processor needs to access data through the bus. The following section will examine why this is so.

Parasitic Capacitance

Functionally, any two conducting bodies separated by an insulating material (including air, plastic, and all nonconductors) form a capacitor. The capacitors we have considered thus far are the type designed and fabricated intentionally for use as components in circuits. In some situations, however, unintentional capacitance may exist in the circuit, in which case it usually is called parasitic capacitance. (Parasitic inductance also is present, but it is usually very small, so we will ignore it.) Consider, for example, the capacitance formed by two parallel wires running side by side on a circuit board. The capacitance of such a two-wire transmission line (Fig. 5-47) is proportional directly to the length of the wires and inversely proportional to a logarithmic function involving d, the spacing between the wires. Thus, C increases with and decreases with d. If the wires are sufficiently long, or sufficiently close to one another, or some combination of the two [as to result in a capacitance of significant magnitude relative to the other capacitances in the circuit] such a wire capacitor (the conductor traces between the different components in the circuit) can slow down the response time of the circuit. In a digital circuit, slower response time means slower switching speed. To explore this subject further, we now examine the impact of parasitic capacitance on the operation of a MOSFET.

5-7.2

CMOS Switching Speed

Recall from Section 4-11 that the gate node in a MOSFET is composed of a metal and a semiconductor separated by a thin layer of silicon dioxide that serves as a dielectric insulator. This geometry is somewhat similar to that of the parallel-plate capacitor of Fig. 5-11. Hence, during normal

“book” — 2015/5/4 — 7:14 — page 307 — #60

5-7 APPLICATION NOTE: PARASITIC CAPACITANCE AND COMPUTER PROCESSOR SPEED

D

CSp

CDn

G

307

VDD

Sp

CGn S

Gp

CSn CGp

CDp PMOS

Dp

(a) NMOS Dn D G CGn

+

n υGS

Gn n iDS

CDn

S

CSn

CDn

+ υin

_

+

NMOS

CGn

Sn

_

υout CSn

_

(a) Original circuit (b) Equivalent circuit

VDD

Figure 5-48: n-channel MOSFET (NMOS): (a) circuit symbol with added parasitic capacitances and (b) equivalent circuit. [In p p a PMOS, parasitic capacitances CD and CS should be shown connected to VDD instead of to ground.]

Sp CDp

operation, the gate (G) and the source (S) nodes form a capacitor between them, as do the gate and the drain (D) nodes. Other parasitic capacitances also exist in a MOSFET, mainly due to charges separated between the source and the large silicon chip and between the drain and the chip. For simplicity, the various parasitic capacitances can be lumped together into an equivalent model containing three capacitances (all connected to ground) from G, S, and D. As shown in Fig. 5-48, these capacitances are designated CGn , CSn , and CDn , respectively, with the superscript “n” denoting that the circuit configuration applies to the n-channel MOSFET (or NMOS for short) whose body node usually is connected to ground. In a p-channel MOSFET, the body node is connected to VDD . Hence, the p p model for PMOS would show parasitic capacitances CD and CS connected to VDD , instead of to ground. Now we are ready to analyze the operation of a CMOS inverter in the presence of parasitic capacitances. The circuit in Fig. 5-49(a) is essentially the same CMOS circuit of Fig. 4-30, except with added parasitic capacitances. The capacitances

G

+ υin

_

D

+

CIN

CDn Sn

υout

_

(b) Simplified circuit Figure 5-49: Common drain inverter circuit with parasitic capacitances. Superscripts “n” and “p” refer to the NMOS and PMOS transistors, respectively.

associated with the n-channel MOSFET are shown connected from terminals Gn , D n , and S n to ground. For the p-channel p MOSFET, capacitance CG is also connected to ground, but for the other two terminals, the capacitances are shown connected

“book” — 2015/5/4 — 7:14 — page 308 — #61

308

CHAPTER 5

(a) Initial condition at t = 0− :

VDD

The capacitances in Fig. 5-50(a) act like open circuits. Also, n = 0 for the NMOS and υin = 0, which means that VGS p VSG = VDD for the PMOS. Under such circumstances,

Sp

+ p VSG

Rs

+ υin

CIN

p iDS

=

G

i3

i1

D

i2

D

n VGS

+

CDn υout

n n iDS = gVGS

_

_

n n = gVGS = 0, iDS

CDp

p gVSG

_ +

p

υout (0− ) = VDD .

(5.145)

υout (0) = VDD .

υout Input

and

Since the voltage across a capacitor cannot change instantaneously,

(a) Equivalent circuit for CMOS inverter υin

p

iDS = gVSG = gVDD , (5.144) where g is the MOSFET gain constant. Furthermore, the PMOS p p behavior is such that, if VSG approaches VDD , the voltage VDS n across the dependent current source goes to zero. With iDS not p conducting and iDS acting like a short circuit, it follows that the voltage across capacitor CDn is

_

Sn

υDD

RC AND RL FIRST-ORDER CIRCUITS

(5.146)

(b) At t ≥ 0: If υin is a step function that changes from 0 to VDD at t = 0, the following pair of responses will take place:

υDD Output

0

t

0

tfall

t

(b) υin(t) and υout(t) Figure 5-50: (a) Equivalent circuit for the CMOS inverter; (b) the response of υout (t) to υin changing states from 0 to VDD at t = 0.

(a) At the input side in the circuit of Fig. 5-50(a), we have an isolated loop comprising υin , Rs , and CIN . In response to the change in υin , capacitor CIN will charge up to a final voltage VDD at a rate governed by the time constant τ = Rs CIN . Through proper choice of Rs (very small), CIN can charge up to VDD so quickly (in comparison with the response time of the output) that it can be assumed that n =V VGS DD immediately after t = 0. n =V (b) At the output side, with VGS DD , it follows that p VSG = 0. Hence,

to VDD . The two MOSFETs share a common gate terminal at the input side and a common drain terminal at the output side. Terminal S n of the NMOS is connected directly to ground, which renders capacitance CSn irrelevant. Terminal S p of the PMOS is connected directly to VDD , which similarly renders p p CS irrelevant. Capacitances CGn and CG both are connected from the common gate terminal to ground and therefore can be combined into an equivalent capacitance CIN . Incorporating these simplifications leads to the circuit shown in Fig. 5-49(b). Our next step is to determine the output response υout (t) to a sudden change of state at the input from υin = 0 to υin = VDD . Let us assume that the change happens at t = 0 and that the circuit was already in a steady-state condition by then.

n = gVDD , iDS

p

p

iDS = gVSG = 0.

and

(5.147)

At node D  , i1 + i2 + i3 = 0,

(5.148)

and at node D, p

n + iDS = gVDD . i3 = iDS

(5.149)

Also, p

i1 = CD

d p d (υout − VDD ) = CD υout , dt dt

(5.150)

and i2 = CDn

d υout . dt

(5.151)

“book” — 2015/5/4 — 7:14 — page 309 — #62

5-7 APPLICATION NOTE: PARASITIC CAPACITANCE AND COMPUTER PROCESSOR SPEED Upon inserting the expressions given by Eqs. (5.149) through (5.151) into Eq. (5.148) and then rearranging terms, we have dυout −gVDD = n p. dt CD + CD

(5.152)

Integrating both sides from 0 to t gives υout |t0 =

−gVDD p CDn + CD

t dt,

(5.153)

0

υout (t) = υout (0) −

υout (t) = VDD 1 −



gVDD p CDn + CD

g p n CD + CD

CDn + CD (0.5 + 0.5) × 10−15 = 10−10 s. = g 10−5

To determine trise , we have to repeat the solution that led to Eq. (5.156) but with υin starting in state 1 (i.e., υin = VDD ) and switching to state 0 at t = 0. Such a process would lead to   g υout (t) = VDD p t. CDn + CD

p

t.

trise = (5.154)

  t .

(5.155)

Plots of υin (t) changing states from 0 to VDD at t = 0 and of the corresponding response υout (t) are displayed in Fig. 5-50(b). We observe that tfall is the time it takes for υout to change states from VDD to zero. From Eq. (5.155), we deduce that p

tfall =

p

tfall =



In view of Eq. (5.146), the expression for υout (t) becomes 

(b) From Eq. (5.156),

The time duration that it takes υout (t) to reach VDD is

which leads to 

309

CDn + CD . g

(5.156)

CDn + CD = tfall . g

Hence, in the presence of parasitic capacitances, Eq. (5.143) is applicable. Namely, fs =

1 1 = = 4.44 GHz. −12 T + 2trise 25 × 10 + 2 × 10−10

In this example, the parasitic capacitances are responsible for slowing down the switching speed of the CMOS processor by about one order of magnitude. In the preceding example, we essentially ignored the input capacitances of the CMOS. Since logic gates are strung along in series such that one gate’s output is the next gate’s input, input capacitances usually are lumped together with the previous gate’s output capacitances. To properly incorporate the roles of both input and output parasitic capacitances, a more thorough treatment is needed than the first-order approximation we carried out in this section. Nevertheless, the approximation did succeed in making the point that at high switching rates parasitic capacitances are important and should not be ignored.

Example 5-19: Processor Speed

The input to a CMOS inverter consists of a sequence of bits, each 25 picoseconds in duration. Determine the maximum switching frequency at which the CMOS inverter can be operated without causing overlap between adjacent bits (pulses) under each of the following conditions: (a) parasitic capacitances totally ignored and (b) parasitic capacitances included. In both cases, p g = 10−5 A/V, and CDn = CD = 0.5 fF. Solution: (a) With T = 25 ps = 25 × 10−12 s and no capacitances to slow down the switching process, the maximum switching frequency is fs =

1 1 = = 40 GHz. T 25 × 10−12

Concept Question 5-26: What is the rationale for adding

parasitic capacitances to nodes G, D, and S in Fig. 5-48? (See ) Concept Question 5-27: What determines the maximum

switching frequency for a CMOS inverter? (See

)

p

n + C = 20 fF Exercise 5-19: A CMOS inverter with CD D

has a fall time of 1 ps. What is the value of its gain constant? Answer: g = 2 × 10−2 A/V. (See

C

)

“book” — 2015/5/4 — 7:14 — page 310 — #63

310

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

t=0 R2 + _

V(3) i

1 kΩ 2.5 V

5 fF

C1 R1

10 kΩ

+ Vout

_ Figure 5-51: RC circuit with an SPST switch.

5-8 Analyzing Circuit Response with Multisim 5-8.1

Modeling Switches in Multisim

Determining the time-dependent behavior of large, complex circuits often is difficult to do and extremely time-consuming. Accordingly, designs of commercial circuits rely heavily on SPICE simulators for evaluating the response of a candidate circuit design before constructing the real version. In this section, we demonstrate how Multisim can be used to analyze the transient response of a circuit driven by a time-dependent source. Because the first-order RC circuit is straightforward to analyze by hand, it makes for a useful example with which we can compare Multisim simulation results to hand calculations. Consider the circuit shown in Fig. 5-51, in which the switch is opened at t = 0 after it had been in the closed position for a long time. Hence, prior to t = 0, the circuit was in a steady state and the capacitor was fully charged with no current flowing through it (behaving like an open circuit). The voltage across the capacitor is designated V(3) (so as to match the Multisim circuit that we will be constructing soon) and is given by V(3) =

2.5 × 10 k = 2.27 V 1 k + 10 k

(@ t = 0− ).

Upon opening the switch, the capacitor will discharge through the 10 k resistor with a time constant given by τdischarge = R1 C1 = 104 × 5 × 10−15 = 50 ps. Likewise, if the switch were to close at a later time after the circuit had fully discharged, the capacitor would again charge up to 2.27 V, but in this case, the time constant would be τcharge = (R1  R2 )C2 =

1 k × 10 k × 5 × 10−15 = 4.54 ps. 11 k

Figure 5-52: Multisim equivalent of the RC circuit in Fig. 5-51.

Thus, the charge-up response of the circuit is much faster (by about one order of magnitude) than its discharge response. To demonstrate the transient behavior of the circuit with Multisim, we construct the circuit model shown in Fig. 5-52 using the component list given in Table 5-6. The only oddity in the circuit is the use of a Voltage-Controlled Switch and a Pulse Generator source to drive it. Multisim does not provide the user the option to use time-programmable switches, so in order to observe the circuit response to multiple opening and closing events of the switch, we use a voltage-controlled switch in combination with an appropriately configured pulse generator. The exact voltage amplitude of the pulse (V2 in Fig. 5-52) is not important (so long as it is larger than the 1 mV threshold of the switch), but the timing of the pulse is critically important, as we want to allow enough time between opening and closing events to observe the complete transient responses of the circuit. Since the longest time constant is 50 ps, double-click on the Pulse Generator and set the Pulse width at 250 ps and the Period at 500 ps so as to provide an adequate time window. Also set the Rise Time and Fall Time to 1 ps. To analyze the behavior, we select Simulate → Analyses → Transient Analysis. Make sure to select an End Time equal to a few periods; 3 ns should suffice. (If you forget this, you may need to abort the simulation to prevent it from running for a long time since the default value is 0.001 s! To abort the simulation or any general Analyses which may be taking too long, go to Simulate → Analyses → Stop Analysis.) In the Output tab, select the non-ground node of the capacitor V(3) and the pulse voltage V(1) for time references. Figure 5-53 shows the

“book” — 2015/5/4 — 7:14 — page 311 — #64

5-8 ANALYZING CIRCUIT RESPONSE WITH MULTISIM

311

Table 5-6: Multisim component list for the circuit in Fig. 5-52. Component

Group

Family

Quantity

Description

1k

Basic

Resistor

1

1 k resistor

10 k

Basic

Resistor

1

10 k resistor

5f

Basic

Capacitor

1

5 fF capacitor

VOLTAGE CONTROLLED SPST

Basic

Switch

1

Switch

DC POWER

Sources

Power Sources

1

2.5 V dc source

PULSE VOLTAGE

Sources

Signal Voltage Source

1

Pulse-generating voltage source

output of the transient analysis. Enabling the Cursor tool in the Grapher window allows the user to read out the exact voltage and time values for any trace.

5-8.2

Modeling Time-Dependent Sources in Multisim

In the previous subsection, we examined how to create switches that toggle with time. What if we wanted to simulate the circuit shown in Fig. 5-54(a) and plot υC over a certain time duration?

The circuit has three time-dependent sources, which would make adding switches and pulse generators rather complicated. Multisim allows us to create the time-dependent sources found in this circuit by using the ABM Voltage and Current sources. In Multisim’s ABM syntax, the step function u(t) is represented by the stp(TIME) function. Also, to guard against Multisim calculating incorrect initial conditions prior to the step function, it is advisable to shift the step-function transition to occur 10 ms after the start of the simulation. Hence, we use the

V(1) V(3)

Figure 5-53: Transient response of the circuit in Fig. 5-52.

“book” — 2015/5/4 — 7:14 — page 312 — #65

312

CHAPTER 5

RC AND RL FIRST-ORDER CIRCUITS

R1 + V1 = 5u(−(t − 0.01)) V _

300 Ω

R2

50 Ω

+ V2 = 3u(t − 0.01) V _

υC

+ C1 _ 100 μF

I1 = 0.1u(t − 0.02) A

(a) Circuit with three time-dependent sources

(b) Multisim circuit

(c) Trace of υC(t) Figure 5-54: Multisim analysis of a circuit containing time-dependent sources.

following ABM expressions: For V1 = 5u(−(t − 0.01)) V:

5*stp(-TIME+0.01)

For V2 = 3u(t − 0.01) V:

3*stp(TIME-0.01)

For I1 = 0.1u(t − 0.02) A:

0.1*stp(TIME-0.02)

Once these expressions have been entered, go to Simulate → Analyses → Transient Analysis. Leave the Start Time at 0 s, and set the End Time to 0.04 s. Under the Output tab, select the voltages V(1), V(2), and V(3) and press Simulate. This generates the plots shown in Fig. 5-54(c).

“book” — 2015/5/4 — 7:14 — page 313 — #66

5-8 ANALYZING CIRCUIT RESPONSE WITH MULTISIM

313

Summary Concepts • The step, ramp, rectangle, and exponential functions can be used to characterize a variety of nonperiodic waveforms. • A capacitor stores electrical energy when a voltage exists across it. • An inductor stores magnetic energy when a current passes through it. • Under dc conditions, a capacitor acts like an open circuit and an inductor acts like a short circuit. • A series RC circuit excited by a dc source exhibits a voltage response (across the capacitor) characterized by an exponential function containing a time constant τ = RC.

• A parallel RL circuit exhibits a current response (through the inductor) that has the same form as the voltage response of the series RC circuit, but for the RL circuit, τ = L/R. • The output voltage of the ideal op-amp RC integrator circuit is directly proportional to the time integral of the input signal. • An integrator circuit becomes a differentiator circuit upon interchanging the locations of R and C. • Parasitic capacitance is often the factor that ultimately limits the processor speed of a computer. • Multisim allows us to evaluate the switching response of a circuit.

Mathematical and Physical Models Unit step  function 0 for t < 0 u(t) = 1 for t > 0

Time-shiftedstep function 0 for t < T u(t − T ) = 1 for t > T

Unit ramp  function 0 for t ≤ 0 r(t) = t for t ≥ 0

Time-shifted  ramp function 0 for t ≤ T r(t − T ) = (t − T ) for t ≥ T

Inductor υ=L

1 i(t) = i(t0 ) + L w= Solenoid

Unit rectangular function (pulse center at t =⎧ T ; pulse length = τ )   ⎪ ⎨0 for t < (T − τ/2), (t − T ) rect = 1 for (T − τ/2) ≤ t ≤ (T + τ/2), ⎪ τ ⎩ 0 for t > (T + τ/2). Capacitor dυ i=C dt t 1 i dt  υ(t) = υ(t0 ) + C w=

1 2

Cυ 2

Parallel plate

t0

(stored electrical energy) εA C= d

di dt

1 2

Li 2 L=

t

υ dt 

t0

(stored magnetic energy) μN 2 S

Series RC circuit response (sudden change at t = 0) υC (t) = υC (∞) + [υ(0) − υ(∞)]e−t/τ τ = RC Parallel RL circuit response (sudden change at t = 0) iL (t) = iL (∞) + [iL (0) − iL (∞)]e−t/τ τ = L/R Op-amp integrator 1 υout (t) = − RC

t

υi dt  + υout (t0 )

t0

Op-amp differentiator dυi υout (t) = −RC dt

“book” — 2015/5/4 — 7:14 — page 314 — #67

314

CHAPTER 5

Important Terms

RC AND RL FIRST-ORDER CIRCUITS

Provide definitions or explain the meaning of the following terms:

air-core solenoid bus bus speed capacitance capacitor charge/discharge charged capacitor circuit response clip clock speed coaxial capacitor dc condition duration of the pulse dynamic circuit early time response electric field electrical permittivity electrical susceptibility equivalent capacitance exponential function ferrite-core inductor

final condition final value first-order circuit first-order RC circuit forced response forcing function inductance initial value iron-core solenoid magnetic field magnetic flux linkage magnetic permeability mica capacitor motherboard mutual inductance nanocapacitor natural decay response natural response negative exponential function nonperiodic waveform

PROBLEMS Section 5-1: Nonperiodic Waveforms 5.1 Generate plots for each of the following step-function waveforms over the time span from −5 to +5 s. (a) υ1 (t) = −6u(t + 3)

op-amp differentiator op-amp integrator parallel-plate capacitor parasitic capacitance periodic waveform permeability permittivity plastic-foil capacitor pulse repetition frequency pulse waveform ramp function RC circuit rectangle function rectangular pulse relative permittivity rise time RL circuit scaling factor self-inductance solenoid source-free

source-free, first-order differential equation static steady-state component steady-state response step function step function response supercapacitor switching frequency (speed) time constant time-shifted ramp function time-shifted step function transient component transient response transmission line uncharged capacitor unit rectangular function unit step function

(b) υ2 (t) = 5r(t + 2) − 5r(t) − 10u(t) (c) υ3 (t) = 10 − 5r(t + 2) + 5r(t)     t +1 t −3 (d) υ4 (t) = 10 rect − 10 rect 2 2     t −1 t −3 (e) υ5 (t) = 5 rect − 5 rect 2 2

(b) υ2 (t) = 10u(t − 4) (c) υ3 (t) = 4u(t + 2) − 4u(t − 2) (d) υ4 (t) = 8u(t − 2) + 2u(t − 4) (e) υ5 (t) = 8u(t − 2) − 2u(t − 4) 5.2 Provide expressions in terms of step functions for the waveforms displayed in Fig. P5.2. *5.3 A 10 V rectangular pulse with a duration of 5 μs starts at t = 2 μs. Provide an expression for the pulse in terms of step functions. 5.4 Generate plots for each of the following functions over the time span from −4 to +4 s. (a) υ1 (t) = 5r(t + 2) − 5r(t) ∗

Answer(s) available in Appendix G.

5.5 Provide expressions for the waveforms displayed in Fig. P5.5 in terms of ramp and step functions. 5.6 Provide plots for the following functions (over a time span and with a time scale that will appropriately display the shape of the associated waveform): (a) υ1 (t) = 100e−2t u(t) (b) υ2 (t) = −10e−0.1t u(t) (c) υ3 (t) = −10e−0.1t u(t − 5) (d) υ4 (t) = 10(1 − e−10 t ) u(t) 3

(e) υ5 (t) = 10e−0.2(t−4) u(t) (f) υ6 (t) = 10e−0.2(t−4) u(t − 4)

“book” — 2015/5/4 — 7:14 — page 315 — #68

PROBLEMS

315

υ1(t)

υ3(t)

υ2(t)

6

6

6

4

4

4

2

2

2

0 1 −2 −1 −2

2

3

4

t (s)

−2 −1 0 1 −2

(a) Step

2

3

t (s)

4

0 1 −2 −1 −2

(b) Bowl

6

6

4

4

4

2

2

2

2

3

4

t (s)

−2 −1 0 1 −2

3

2

t (s)

4

0 1 −2 −1 −2

(e) Hat

(d) Staircase down

4

2

3

4

(f) Square wave

Figure P5.2: Waveforms for Problem 5.2.

υ1(t)

υ2(t)

4

4

2 −2

2 0

−2

t (s) 2

4

6

−2

0

t (s) 2

4

−4 (a) “Vee”

(b) Mesa υ3(t) 4 2

−2

0 −2

t (s)

υ6(t)

6

0 1 −2 −1 −2

3

(c) Staircase up

υ5(t)

υ4(t)

2

t (s) 2

4

6

−4 (c) Sawtooth Figure P5.5: Waveforms for Problem 5.5.

6

t (s)

“book” — 2015/5/4 — 7:14 — page 316 — #69

316

CHAPTER 5

*5.7 After opening a certain switch at t = 0 in a circuit containing a capacitor, the voltage across the capacitor started decaying exponentially with time. Measurements indicate that the voltage was 7.28 V at t = 1 s and 0.6 V at t = 6 s. Determine the initial voltage at t = 0 and the time constant of the voltage waveform.

RC AND RL FIRST-ORDER CIRCUITS

(d) What is the maximum amount of energy stored in the capacitor, and when does it occur? 5.11 Suppose the waveform shown in Fig. P5.10 is the current i(t) through a 0.2 mF capacitor (rather than the voltage) and its peak value is 100 μA. given that the initial voltage on the capacitor was zero at t = −4 s, determine and plot υ(t). 5.12 The current through a 40 μF capacitor is given by a rectangular pulse as

Section 5-2: Capacitors



5.8 After plotting the voltage waveform, obtain expressions and generate plots for i(t), p(t), and w(t) for a 0.2 mF capacitor. The voltage waveforms are given by (a) υ1 (t) = 5r(t) − 5r(t − 2) V (b) υ2 (t) = 10u(−t) + 10u(t) − 5r(t − 2) + 5r(t − 4) V (c) υ3 (t) = 15u(−t) + 15e−0.5t u(t) V (d) υ4 (t) = 15[1 − e−0.5t ] u(t) V *5.9 In response to a change introduced by a switch at t = 0, the current flowing through a 100 μF capacitor, defined in accordance with the passive sign convention, was observed to be i(t) = −0.4e−0.5t mA

i(t) = 40 rect

t −1 2

 mA.

If the capacitor was initially uncharged, determine υ(t), p(t), and w(t). 5.13 The voltage across a 0.2 mF capacitor was 20 V until a switch in the circuit was opened at t = 0, causing the voltage to vary with time as υ(t) = (60 − 40e−5t ) V

(for t > 0).

(a) Did the switch action result in an instantaneous change in υ(t)?

(for t > 0).

If the final energy stored in the capacitor (at t = ∞) is 0.2 mJ, determine υ(t) for t ≥ 0.

(b) Did the switch action result in an instantaneous change in the current i(t)?

5.10 The voltage υ(t) across a 20 μF capacitor is given by the waveform shown in Fig. P5.10.

(c) How much initial energy was stored in the capacitor at t = 0?

υ (V)

5.14 Determine voltages υ1 to υ4 in the circuit of Fig. P5.14 under dc conditions.

100

−4

−2

0

(d) How much final energy will be stored in the capacitor (at t = ∞)?

2

4

+

t (s)

Figure P5.10: Waveform for Problems 5.10 and 5.11.

(b) Specify the time interval(s) during which power transfers into the capacitor and that (those) during which it transfers out of the capacitor. (c) At what instant in time is the power transfer into the capacitor a maximum? And at what instant is the power transfer out of the capacitor a maximum?

C1

+ _υ1

30 kΩ C2

_

C3

20 kΩ

(a) Determine and plot the corresponding current i(t).

υ3

+ _ υ2

15 kΩ 5 kΩ C4

+ _ 15 V

+ _υ4

Figure P5.14: Circuit for Problem 5.14.

10 kΩ

“book” — 2015/5/4 — 7:14 — page 317 — #70

PROBLEMS

317

*5.15 Determine voltages υ1 to υ3 in the circuit of Fig. P5.15 under dc conditions.

5.18 Reduce the circuit in Fig. P5.18 into a single equivalent capacitor at terminals (a, b). Assume that all initial voltages are zero at t = 0.



+ 10 Ω

18 Ω

+ 40 V _



υ2

_

+

υ3

_

60 μF

20 μF

C



+

υ1

C C

_

Figure P5.15: Circuit for Problem 5.15.

10 V + _

3 kΩ

+ _υ1

20 μF

_ + 2V

υ2

+ _

C C

Figure P5.18: Circuit for Problem 5.18.

40 μF a

5F

3F 6F

Figure P5.16: Circuit for Problem 5.16.

*5.17 Reduce the circuit in Fig. P5.17 into a single equivalent capacitor at terminals (a, b). Assume that all initial voltages are zero at t = 0.

6 μF

5F

c

6F d

b 3F

8 μF

C

*5.19 For the circuit in Fig. P5.19, find Ceq at terminals (a, b). Assume all initial voltages to be zero.

20 kΩ 3 kΩ

10 kΩ

b

C

5.16 Determine the voltages across the two capacitors in the circuit of Fig. P5.16 under dc conditions.

40 kΩ

a

C

10 μF

40 kΩ

C

5F

Figure P5.19: Circuit for Problems 5.19 and 5.20.

5.20

Find Ceq at terminals (c, d) in the circuit of Fig. P5.19.

a 3 μF

6 μF

12 μF

b 10 μF Figure P5.17: Circuit for Problems 5.17 and 5.21.

*5.21 Assume that a 120 V dc source is connected at terminals (a, b) to the circuit in Fig. P5.17. Determine the voltages across all capacitors. 5.22 Determine (a) the amount of energy stored in each of the three capacitors shown in Fig. P5.22, (b) the equivalent capacitance at terminals (a, b), and (c) the amount of energy stored in the equivalent capacitor.

“book” — 2015/5/4 — 7:14 — page 318 — #71

318

CHAPTER 5

5.26 The waveform shown in Fig. P5.26 represents the voltage across a 0.2 H inductor for t ≥ 0. If the current flowing through the inductor is −20 mA at t = 0, determine the current i(t) for t ≥ 0.

20 μF

a 10 kΩ

6 μF

+ 15 V _

RC AND RL FIRST-ORDER CIRCUITS

5 μF

b

υ (mV)

Figure P5.22: Circuit for Problem 5.22.

20 10

Section 5-3: Inductors

0

5.23 After plotting the current waveform, obtain expressions and generate plots for υ(t), p(t), and w(t) for a 0.5 mH inductor. The current waveforms are given by (a) i1 (t) = 0.2r(t − 2) − 0.2r(t − 4) − 0.2r(t − 8) + 0.2r(t − 10) A (b) i2 (t) =

2u(−t) + 2e−0.4t

0

2

t (s)

3

Figure P5.26: Voltage waveform for Problem 5.26.

5.27 The waveform shown in Fig. P5.27 represents the voltage across a 50 mH inductor. Determine the corresponding current waveform. Assume i(0) = 0.

u(t) A

(c) i3 (t) = −4(1 − e−0.4t ) u(t) A 5.24 The current i(t) passing through a 0.1 mH inductor is given by the waveform shown in Fig. P5.24.

υ

(a) Determine and plot the corresponding voltage υ(t) across the inductor.

10 cos (πt/4) (mV)

10 mV

(b) Specify the time interval(s) during which power is transferred into the inductor and that (those) during which power transfers out of the inductor. Also specify the amount of energy transferred in each case.

0

2

4

t (s)

−10 mV i (A) Figure P5.27: Voltage waveform for Problem 5.27.

3 t (s) −4

−2

0

2

4

5.28 For the circuit in Fig. P5.28, determine the voltage across C and the currents through L1 and L2 under dc conditions.

Figure P5.24: Current waveform for Problem 5.24.

L1 = 2 mH *5.25 Activation of a switch at t = 0 in a certain circuit caused the voltage across a 20 mH inductor to exhibit the voltage response υ(t) = 4e−0.2t mV

10 Ω 15 Ω

2A



L2 = 4 mH

C = 20 μF

(for t ≥ 0).

Determine i(t) for t ≥ 0 given that the energy stored in the inductor at t = ∞ is 0.64 mJ.

Figure P5.28: Circuit for Problem 5.28.

“book” — 2015/5/4 — 7:14 — page 319 — #72

PROBLEMS

319

*5.29 For the circuit in Fig. P5.29, determine the voltages across C1 and C2 and the currents through L1 and L2 under dc conditions.

3 3

4 C2 = 2 μF

a

3

1 4

Leq

10 Ω 5Ω

L1 = 2 H



L2 = 6 H

1

3

3

b



+ _ 30 V

C1 = 1 μF

3 3

Figure P5.32: Circuit for Problem 5.32. Figure P5.29: Circuit for Problem 5.29.

Section 5-4: Response of the RC Circuit 5.30 All elements in Fig. P5.30 are 10 mH inductors. Determine Leq .

L

L

Leq

L

L

5.33 After having been in position 1 for a long time, the switch in the circuit of Fig. P5.33 was moved to position 2 at t = 0. Given that V0 = 12 V, R1 = 30 k , R2 = 120 k , R3 = 60 k , and C = 100 μF, determine: (a) iC (0− ) and υC (0− )

L

(b) iC (0) and υC (0) (c) c iC (∞) and υC (∞) (d) υC (t) for t ≥ 0

L

L

(e) iC (t) for t ≥ 0

Figure P5.30: Circuit for Problem 5.30.

R1 *5.31 The values of all inductors in the circuit of Fig. P5.31 are in millihenrys. Determine Leq .

a

3

5

2

iC

i1

+ V0 _

1 R2

R3

C

υC

8 Figure P5.33: Circuit for Problems 5.33 and 5.34.

Leq 4

8

12 6

6

12

b Figure P5.31: Circuit for Problem 5.31.

5.32 Determine Leq at terminals (a, b) in the circuit of Fig. P5.32. All inductor values are in millihenrys.

5.34 Repeat Problem 5.33, but with the switch having been in position 2 for a long time, and then moved to position 1 at t = 0. 5.35 The circuit in Fig. P5.35 contains two switches, both of which had been open for a long time before t = 0. Switch 1 closes at t = 0, and switch 2 follows suit at t = 5 s. Determine and plot υC (t) for t ≥ 0 given that V0 = 24 V, R1 = R2 = 16 k , and C = 250 μF. Assume υC (0) = 0.

“book” — 2015/5/4 — 7:14 — page 320 — #73

320

CHAPTER 5

Switch 1

Switch 2

t=0

υC

C

t=0

R1

R1 + V0 _

RC AND RL FIRST-ORDER CIRCUITS

t=5s

R2

Vs

i

+ _

C1

*5.36 The circuit in Fig. P5.36 was in steady state until the switch was moved from terminal 1 to terminal 2 at t = 0. Determine υ(t) for t ≥ 0 given that I0 = 21 mA, R1 = 2 k , R2 = 3 k , R3 = 4 k , and C = 50 μF.

*5.39 The switch in the circuit of Fig. P5.39 had been in position 1 for a long time until it was moved to position 2 at t = 0. Determine υ(t) for t ≥ 0, given that I0 = 6 mA, V0 = 18 V, R1 = R2 = 4 k , and C = 200 μF.

υ

I0

R1

2

υ

1

C t=0

C2

Figure P5.38: Circuit for Problem 5.38.

Figure P5.35: Circuit for Problem 5.35.

1

R2

R1

I0

C

2

+ _ V0

R2

R2

R3 Figure P5.39: Circuit for Problems 5.39 and 5.40.

Figure P5.36: Circuit for Problem 5.36.

5.37 Prior to t = 0, capacitor C1 in the circuit of Fig. P5.37 was uncharged. For I0 = 5 mA, R1 = 2 k , R2 = 50 k , C1 = 3 μ F, and C2 = 6 μ F, determine: (a) The equivalent circuit involving the capacitors for t ≥ 0. Specify υ1 (0) and υ2 (0). (b) i(t) for t ≥ 0. (c) υ1 (t) and υ2 (t) for t ≥ 0.

R2

1 I0

R1

2 C1

t=0 υ1

5.40 Repeat Problem 5.39, but reverse the switching sequence. [Switch starts in position 2 and is moved to position 1 at t = 0.] 5.41 Determine i(t) for t ≥ 0 where i is the current passing through R3 in the circuit of Fig. P5.41. The element values are υs = 16 V, R1 = R2 = 2 k , R3 = 4 k , and C = 25 μF. Assume that the switch had been open for a long time prior to t = 0.

i C2

t=0

R1 υ2

Figure P5.37: Circuit for Problem 5.37.

5.38 The switch in the circuit of Fig. P5.38 had been closed for a long time before it was opened at t = 0. Given that Vs = 10 V, R1 = 20 k , R2 = 100 k , C1 = 6 μF, and C2 = 12 μF, determine i(t) for t ≥ 0.

υs

+ _

C

υ

R2

i R3

Figure P5.41: Circuit for Problems 5.41 to 5.43.

5.42 Repeat Problem 5.41, but start with the switch being closed prior to t = 0 and then opened at t = 0. *5.43 Consider the circuit in Fig. P5.41, but without the switch. If the source υs represents a 12 V, 100 ms long rectangular

“book” — 2015/5/4 — 7:14 — page 321 — #74

PROBLEMS

321

pulse that starts at t = 0 and the element values are R1 = 6 k , R2 = 2 k , R3 = 4 k , and C = 15 μF, determine the voltage response υ(t) for t ≥ 0. 5.44 Given that in Fig. P5.44, I1 = 4 mA, I2 = 6 mA, R1 = 3 k , R2 = 6 k , and C = 0.2 mF, determine υ(t). Assume the switch was connected to terminal 1 for a long time before it was moved to terminal 2.

R1

2

+ V0 _

iL 1

R3

R2

L

υL

Figure P5.46: Circuit for Problems 5.46 and 5.47.

1

2 t=0

I1

R1

R2 C

I2

υ

5.47 Repeat Problem 5.46, but with the switch having been in position 2 for a long time and then moved to position 1 at t = 0. *5.48 Determine i(t) for t ≥ 0 given that the circuit in Fig. P5.48 had been in steady state for a long time prior to t = 0. Also, I0 = 5 A, R1 = 2 , R2 = 10 , R3 = 3 , R4 = 7 , and L = 0.15 H.

Figure P5.44: Circuit for Problem 5.44.

*5.45 Determine υC (t) in the circuit of Fig. P5.45 for t ≥ 0, given that the switch had been closed for a long time prior to t = 0.

i L

1 1 kΩ

t=0 1 kΩ

20 V

+ _

2 kΩ

+

υC

I0

R1

_

1 kΩ

Figure P5.45: Circuit for Problem 5.45.

5.49 For the circuit in Fig. P5.49, determine iL (t) and plot it as a function of t for t ≥ 0. The element values are I0 = 4 A, R1 = 6 , R2 = 12 , and L = 2 H.Assume that iL = 0 before t = 0.

Section 5-5: Response of the RL Circuit 5.46 After having been in position 1 for a long time, the switch in the circuit of Fig. P5.46 was moved to position 2 at t = 0. Given that V0 = 12 V, R1 = 30 , R2 = 120 , R3 = 60 , and L = 0.2 H, determine: (a) iL

(0− )

and υL

R3

R4

Figure P5.48: Circuit for Problem 5.48.

10 μF 2 kΩ

R2

t=0

2

I0

R1

t=0

L

t = 0.5 s

R2

Figure P5.49: Circuit for Problem 5.49.

(0− )

(b) iL (0) and υL (0) (c) iL (∞) and υL (∞) (d) iL (t) for t ≥ 0 (e) υL (t) for t ≥ 0

*5.50 After having been in position 1 for a long time, the switch in the circuit of Fig. P5.50 was moved to position 2 at t = 0. Determine i1 (t) and i2 (t) for t ≥ 0, given that I0 = 6 mA, R0 = 12 , R1 = 10 , R2 = 40 , L1 = 1 H, and L2 = 2 H.

“book” — 2015/5/4 — 7:14 — page 322 — #75

322

CHAPTER 5

t=0

2 I0

i1

i2

L1

R0

R1

R2

1

RC AND RL FIRST-ORDER CIRCUITS

L2

R3 i

+ υs(t) _

R2

R1

L

(a) Circuit υs(t)

Figure P5.50: Circuit for Problem 5.50.

12 V 5.51 Derive an expression for i2 (t) in the circuit of Fig. P5.51 in terms of the circuit variables, given that Is is a dc current source and the switch was closed at t = 0 after it had been open for a long time.

t

0 (b) υs(t) for Problem 5.53 υs(t)

R1 t=0 Rs

Is

R2

i2

12 V

L 0

t (s)

3 (c) υs(t) for Problem 5.54

Figure P5.51: Circuit for Problem 5.51.

υs(t) 5.52

Determine iL (t) in the circuit of Fig. P5.52 for t ≥ 0.

0.4 Vx +_

10 Ω

0 t=0

+ 1A

V

25 Ω

_x

12 V

π/6

π/3

π/2

t (s)

iL 5H

(d) υs(t) for Problem 5.55

5Ω Figure P5.53: Circuit and excitation voltages for Problems Figure P5.52: Circuit for Problem 5.52.

*5.53 In the circuit of Fig. P5.53(a), R1 = R2 = 20 , R3 = 10 , and L = 2.5 H. Determine i(t) for t ≥ 0 given that υs (t) is the step function described in Fig. P5.53(b).

5.53 to 5.55.

5.54 Repeat Problem 5.53 for the triangular-source excitation given in Fig. P5.53(c).  eax Hint : xeax dx = 2 (ax − 1). a

“book” — 2015/5/4 — 7:14 — page 323 — #76

PROBLEMS

323

5.55 Repeat Problem 5.53 for the sinusoidal-source excitation υs (t) = 12 sin 6t V displayed in Fig. P5.53(d).

υi 12 V

 Hint :

eax sin bx dx = eax

[a sin bx − b cos(bx)] . a 2 + b2

0 2

4

6

8 10 12

t (s)

-12 V (a) Waveform of υi(t)

*5.56 The switch in the circuit of Fig. P5.56 was moved from position 1 to position 2 at t = 0, after it had been in position 1 for a long time. If L = 80 mH, determine i(t) for t ≥ 0.

50 kΩ

2 μF 1

_ +

υi

2

Vcc = 6 V

t=0 10 mA

20 Ω

i

υout

40 Ω

(b) Op-amp circuit

20 mA

L

Figure P5.59: Waveform and circuit for Problem 5.59.

Figure P5.56: Circuit for Problems 5.56 and 5.57.

*5.60

Relate υout to υi in the circuit of Fig. P5.60.

+ _

υi 5.57 Repeat Problem 5.56, but with the switch having been in position 2 and then moved to position 1 at t = 0.

υout C

R

5.58 Determine i(t) for t ≥ 0 due to the rectangular-pulse excitation in the circuit of Fig. P5.58.

Figure P5.60: Circuit for Problem 5.60.

6Ω i + _

16 V 0

8 mH

12 Ω

5.61 Develop the relationship between the output voltage υout and the input voltage υi for the circuit in Fig. P5.61.

4 ms Figure P5.58: Circuit for Problem 5.58.

R υi

R

_ C

+

υout

Section 5-6: RC Op-Amp Circuits Figure P5.61: Circuit for Problem 5.61.

5.59 The input-voltage waveform shown in Fig. P5.59(a) is applied to the circuit in Fig. P5.59(b). Determine and plot the corresponding υout (t).

5.62 Relate υout to υi in the circuit of Fig. P5.62. Assume υC = 0 at t = 0.

“book” — 2015/5/4 — 7:14 — page 324 — #77

324

CHAPTER 5

C R1

υi

R2

C

_

*5.63 Relate iout (t) to υi (t) in the circuit of Fig. P5.63. Evaluate it for υC (0) = 3 V, R = 10 k , C = 50 μF, and υi (t) = 9 u(t) V.

+ _

R2

_ +

υout

Figure P5.65: Circuit for Problem 5.65.

5.66 Design a single op-amp circuit with a 40 μF capacitor to generate a circuit output given by t

C iout

υC

R3

υs(t) = Au(t)

Figure P5.62: Circuit for Problem 5.62.

Vcc = 12 V

R1

υout

+

υi

RC AND RL FIRST-ORDER CIRCUITS

R

υout (t) =

[6 − 2υs (t  )] dt  = 6t − 2

0

t

υs (t  ) dt 

(V),

0

where υs (t) is any input voltage source that starts at t = 0. 5.67 Design a circuit that can perform the following relationship between its output and input voltages:

Figure P5.63: Circuit for Problem 5.63.

t 5.64

υout = −100

Determine υout (t) in the circuit of Fig. P5.64 for t ≥ 0.

1 kΩ + _

υi dt, 0

1 mF

with υout (0) = 0 at t = 0. You are limited to one op-amp, one capacitor that does not exceed 0.1 F, and any resistor(s) of your choice.

2 kΩ

5.68 The two-stage op-amp circuit in Fig. P5.68 is driven by an input step voltage given by υi (t) = 10 u(t) mV. If Vcc = 10 V for both op amps and the two capacitors had no charge prior to t = 0, determine and plot: *(a) υout1 (t) for t ≥ 0; (b) υout2 (t) for t ≥ 0.

_ +

12u(t) V

υout 5 kΩ

4 μF Figure P5.64: Circuit for Problem 5.64.

5.65

υi 5 kΩ

+

In the circuit of Fig. P5.65:

(a) Derive an expression for υout (t) for t ≥ 0 in terms of R1 , R2 , R3 , C, and A. *(b) Evaluate the expression for R1 = 1 k , R3 = 2 k , C = 0.25 mF, and A = 12 V.

_

R2 = 5 k ,

5 μF υout1 1 MΩ Vcc = 10 V

_ +

υout2 Vcc = 10 V

Figure P5.68: Op-amp circuit for Problem 5.68.

“book” — 2015/5/4 — 7:14 — page 325 — #78

PROBLEMS

325

υ

R

C

R

_

R

+

υ1

R

_

R

+

υ2

_

R

+

υ3

2R R

υs

_ +

υ4

Figure P5.71: Circuit for Problem 5.71. 5.69 Design a single op-amp circuit that can perform the operation t υout = − (5υ1 + 2υ2 + υ3 ) dt. 0

5.70 Design a single op-amp circuit that can perform the operation t  υ1 υ2 υ3  + + dt. iout = − 100 200 400 0

5.71 Show that the op-amp circuit in Fig. P5.71 (in which R = 10 k and C = 20 μF) simulates the differential equation dυ + 5υ = 10υs . dt 5.72 Design an op-amp circuit that can solve the differential equation dυ + 0.2υ = 4 sin 10t dt with υ(0) = 0. Hint: See Problem 5.71.

Sections 5-7 and 5-8: Parasitic Capacitance and Multisim Analysis *5.73 In real transistors, both the MOSFET gain g and parasitic p capacitances CDn and CD depend on the size of the transistor. Assuming the functional relationships g = 106 W

and

where W is the transistor width in meters, how small should W be in order for the CMOS inverter to have a fall time of 1 ns? [The width of modern digital MOSFETs varies between 40 nm and 4 μm.] 5.74 Draw and simulate in Multisim the circuit in Fig. 5-43(a) of Example 5-15. Using the Grapher tool, plot υout (t) for t ≥ 0. 5.75 Consider the circuit in Fig. P5.75. Switch S1 begins in the closed position and opens at t = 0. Switch S2 begins in the open position and toggles between the open and closed positions every 250 ps. Model this circuit in Multisim and plot υ0 and υ1 as a function of time until all nodes are discharged below 1 mV.

t=0

S2 1 kΩ

+ 2.5 V _

p

CDn = CD = (2.5 × 103 )W 2 ,

5.5 kΩ

S1 5 fF

10 kΩ

υ0

Figure P5.75: Circuit for Problem 5.75.

20 fF

υ1

“book” — 2015/5/4 — 7:14 — page 326 — #79

326

CHAPTER 5

R1

R1

R1

C1

C1

R1

RC AND RL FIRST-ORDER CIRCUITS

R1

C1

+

C1

υout1

C1

_

R2 + υs(t) _

R2

R2

C2

C2

R2

R2

C2

+

C2

υout2

C2

_

Figure P5.76: Circuit for Problem 5.76 with R1 = R2 = 10 , C1 = 7 pF, and C2 = 5 pF.

5.76 A step voltage source υs (t) sends a signal down two transmission lines simultaneously (Fig. P5.76). In Multisim, the step voltage may be modeled as a 1 V square wave with a period of 10 ns. Model the circuit in Multisim and answer the following questions:

+

t=0

υc

+

1V

_

υa 5 fF

Ω M 1



(b) By how much? Hint: When using cursors in the Grapher View, select a trace, then right-click on a cursor and select Set Y Value, and enter 750 m. This will give you the exact time point at which that trace equals 0.75 V.

_

5 3.

(a) If a detector registers a signal when the output voltage reaches 0.75 V, which signal arrives first?

1 fF

10 kΩ

_

5.77 Consider the delta topology in Fig. P5.77. Use Multisim to generate response curves for υa , υb , and υc . Apply Transient Analysis with TSTOP = 3 × 10−10 s.

2 fF υb

+

Figure P5.77: Circuit for Problem 5.77.

5.78 Use Multisim to generate a plot for current i(t) in the circuit in Fig. P5.78 from 0 to 15 ms. (a) υin (t) is a 0-to-1 V square wave with a period of 1 ms and a 50 percent duty cycle. Plot the output from 0 to 10 ms.

5.79 Construct the integrator circuit shown in Fig. P5.79, using a 3-terminal virtual op amp. Print the output corresponding to each of the following input signals:

(b) υin (t) = −0.2t V. Plot the output from 0 to 50 ms.

L1

R1 υs(t) = [−5u(-t) + 5u(0.003 − t)] V

+ _

90 Ω

i

500 mH

R2

Figure P5.78: Circuit for Problem 5.78.

220 Ω

0.1u(t − 0.003) A

“book” — 2015/5/4 — 7:14 — page 327 — #80

PROBLEMS

327

υ(t)

C1

υin

R1 100 Ω

100 μF

_

+

8e−(t − 0.02)/0.008 V for t > 0.02 s

8V υout

Figure P5.79: Circuit for Problem 5.79.

0

10

20

30

40

50

t (ms)

Figure m5.1 Voltage waveform for Problem m5.1.

Potpourri Questions 5.80 Calculate the plate area required to store 1 MJ of energy in a traditional air-filled parallel plate capacitor at a voltage of 10 V. Assume the plate separation to be 1 cm.

(b) Determine the time at which the inductor current reaches its maximum value.

5.81 What are the advantages and disadvantages of supercapacitors relative to a lithium-ion battery?

(c) Calculate the total peak-to-peak range of inductor current; i.e., the maximum value minus the minimum value.

5.82 Is the memory stored on a hard disk drive volatile or nonvolatile? What is the advantage of perpendicular magnetic recording over the standard recording method? 5.83 How does the proximity of a finger change the capacitance of a pixel in a touchscreen? How does the MEMS capacitor measure the acceleration of a moving vehicle?

υ(t) 9V

0

0.1

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m5.1 Capacitors: The voltage υ(t) across a 10 μF capacitor is given by the waveform shown in Fig. m5.1.

0.2

0.3

0.4

t (ms)

−9 V

Figure m5.2 Voltage waveform for Problem m5.2.

(b) Calculate the values of the capacitor current at times 0, 25, and 30 ms.

m5.3 Response of the RC Circuit: Figure m5.3(a) shows a resistor-capacitor circuit with a pair of switches and Fig. m5.3(b) shows the switch opening-closing behavior as a function of time. The initial capacitor voltage is −9 V. Component values are R1 = 10 k , R2 = 3.3 k , R3 = 2.2 k , C = 1.0 μF, V1 = 9 V, and V2 = −15 V.

m5.2 Inductors: The voltage υ(t) across a 33 mH inductor is given by the sinusoidal pulse waveform shown in Fig. m5.2.

(a) Determine the equation that describes υ(t) over the time range 0 to 50 ms.

(a) Determine the equation for the capacitor current i(t) and plot it over the time period from 0 to 50 ms.

(a) Determine the equation for the inductor current i(t) and plot it over the time period from 0 to 0.4 ms. Assume zero initial inductor current.

(b) Plot υ(t) over the time range 0 to 50 ms. (c) Determine the values of υ(t) at the times 5, 15, 25, 35, and 45 ms.

“book” — 2015/5/4 — 7:14 — page 328 — #81

328

CHAPTER 5

R1 V1

Sw1

Rs

R3

Sw2

+

+ _

υ(t)

_

+ _ V2

R2

C

RC AND RL FIRST-ORDER CIRCUITS

Vbatt

Rw

+ _

+ υ(t)

L

_

Rload

(a) Circuit Sw1

Figure m5.4 Circuit for Problem m5.4.

Closed

Open

10

20

30

40

t (ms)

50

(a) Determine the load voltage υ after the switch had been closed for a long time.

Sw2

(b) Determine the equation that describes υ(t) after the switch opens at time t = 0.

Closed

Open

Resistor Rs models the finite resistance of an electronic analog switch and Rw models the finite winding resistance of the inductor. Component values are: Rs = 16 , Rw = 90 , Rload = 680 , L = 33 mH, and Vbatt = 1.5 V.

10

20

30

40

t (ms)

50

(c) Determine the magnitude of the peak value of υ(t). How many times larger is this value compared to the battery voltage Vbatt ? (d) State the value of the circuit time constant τ with the switch open. Plot υ(t) over the time range −τ ≤ t ≤ 5τ .

(b) Figure m5.3 Voltage waveform for Problem m5.3.

m5.4 Response of the RL Circuit: The circuit of Fig. m5.4 demonstrates how an inductor can produce a high-voltage pulse across a load resistor Rload that is considerably higher than the circuit’s power supply Vbatt , a 1.5 V “AA” battery. Highvoltage pulses drive photo flash bulbs, strobe lights, and cardiac defibrillators, as examples.

m5.5 RC Differentiator: The circuit in Fig. m5.5 is a differentiator. Find υout (t), given that υs (t) is a 300 Hz sinusoid with an amplitude of 3 V. You will need to use the myDAQ’s Function Generator and Oscilloscope for this problem. m5.6 RC Integrator: The circuit in Fig. m5.6 is an RC integrator circuit. Find υout (t), given that υs (t) is a 100 Hz sinusoid with an amplitude of 5 V. You will need to use the myDAQ’s Function Generator and Oscilloscope for this problem.

1 kΩ C1

~

υs(t) = 3 cos(600πt) V + −

1 μF

_ +

R1 10 kΩ

Figure m5.5 A differentiator circuit.

+ υout(t) _

“book” — 2015/5/4 — 7:14 — page 329 — #82

PROBLEMS

329

C1 1 μF R1

~

υs(t) = 5 cos(200πt) V + −

100 kΩ

_ +

R2 1 kΩ

Figure m5.6 Circuit for Problem m5.6.

+ υout(t) _

“book” — 2015/5/4 — 7:16 — page 330 — #1

6

CHAPTER

RLC Circuits

Contents 6-1 6-2 TB15 6-3 6-4 6-5 6-6 6-7 TB16 6-8 TB17 6-9

Overview, 331 Initial and Final Conditions, 331 Introducing the Series RLC Circuit, 334 Micromechanical Sensors and Actuators, 337 Series RLC Overdamped Response (a > ω0 ), 341 Series RLC Critically Damped Response (a = ω0 ). 346 Series RLC Underdamped Response (a < ω0 ), 348 Summary of the Series RLC Circuit Response, 349 The Parallel RLC Circuit, 353 RFID Tags and Antenna Design, 356 General Solution for Any Second-Order Circuit with dc Source, 359 Neural Stimulation and Recording, 363 Multisim Analysis of Circuits Response, 369 Summary, 373 Problems, 374

Objectives Learn to: 

Analyze series and parallel RLC circuits containing dc sources and switches.



Analyze RC op-amp circuits. Understand RFID circuits.



To receiver circuits R T υout(t)

~+− υ

s

RFID transceiver

Ls

Magnetic field

Lp C p

υC

RFID tag

Rp

“book” — 2015/5/4 — 7:16 — page 331 — #2

6-1

INITIAL AND FINAL CONDITIONS

Overview In this chapter we evaluate the operation of second-order RLC circuits—those with any combination of two inductors and/or capacitors—in response to dc sources (the response of RLC circuits to ac sources is covered in Chapter 7). These circuits are particularly interesting because they allow us to design oscillators and resonators for communication and wireless power transmission systems, or to create sensors that use the oscillation or resonance to detect capacitive (usually) or inductive (rarely) changes caused by environmental parameters (moisture, pressure, proximity, etc.). One particularly interesting example is wireless power transfer for radiofrequency ID (RFID) systems, as described in Section 6-9 and Technology Brief 16. Using two inductors and a capacitor, the current in one loop is converted into voltage in the capacitor, that can then be used to power the RFID circuit. The currents and voltages of the first-order RC and RL circuits we examined in the preceding chapter were characterized by first-order differential equations. A key provision of a first-order circuit is that it is reducible to a single series or parallel circuit containing a single capacitor or a single inductor, in addition to sources and resistors. If a circuit contains two capacitors, as in Fig. 6-1(a), and if the circuit architecture is such that it is not possible to combine the two capacitors into a single in-series or in-parallel equivalent, then the circuit does not qualify as a first-order circuit. The two-capacitor circuit is a second-order circuit characterized by a second-order differential equation. The same is true for the two-inductor circuit in part (b) and for the series and parallel RLC circuits shown in parts (c) and (d) of the same figure.  A second-order circuit may contain any combination of two energy-storage elements (2 capacitors, 2 inductors, or one of each), provided like-elements cannot be replaced with a single-element equivalent.  In general, the order of a circuit, and hence the order of the differential equation describing any of its currents or voltages, is governed by the number of irreducible storage elements (capacitors and inductors) contained in the circuit. The complexity of the solution depends on the order of the differential equation and the character of the excitation source. In this chapter we examine the response of series and parallel RLC circuits to dc excitations, and we do so by solving their differential equations in the time domain. Time-domain solutions are reasonably tractable, so long as the forcing function is a dc source or a rectangular pulse, and the differential equation describing the voltages and currents in the circuit is

331 not higher than second order. For more complicated circuits, a more robust method of solution is called for, such as the Laplace transform analysis technique introduced in Chapter 12, which is perfectly suited to deal with a wide range of circuits and any type of realistic forcing function, including pulses and sinusoids.

6-1

Initial and Final Conditions

The general form of the solution of the differential equation associated with a second-order circuit always includes a number of unknown constants. To determine the values of these constants, we usually match the solution to known values of the voltage or current under consideration. For a circuit where the solution we seek is for the time period following a sudden change (such as when a SPST switch is closed or opened, or when a SPDT switch is moved from one terminal to another)

R1 υs

+ _

t=0

R2 C1

(a)

C2

2 capacitors R1

υs

+ _

t=0

R2 L1

(b)

L2

2 inductors L

R υs

+ _

t=0

C

(c)

is (d)

Series RLC t=0

R

L

C

Parallel RLC

Figure 6-1: Examples of second-order circuits.

“book” — 2015/5/4 — 7:16 — page 332 — #3

332

CHAPTER 6

we can analyze the circuit conditions at the beginning and at the end of that time period and then use the results to match the solution of the differential equation. We call the process invoking initial and final conditions. Analyzing a circuit in its initial and final states relies on the following fundamental properties:

R2 iL

R1 Vs

+ + __

Example 6-1: Initial and Final Values

The circuit in Fig. 6-2(a) contains dc source Vs and a switch that had been in position 1 for a long time prior to t = 0. Determine: (a) initial values υC (0) and iL (0), (b) iC (0) and υL (0), and (c) final values υC (∞) and iL (∞). Solution: (a) To determine υC (0) and iL (0), we analyze the circuit configuration at t = 0− (before moving the switch), whereas to determine iC (0) and υL (0), we analyze the circuit configuration at t = 0 (after moving the switch). At t = 0− , the circuit is equivalent to the arrangement shown in Fig. 6-2(b), in which C has been replaced with an open circuit and L with a short circuit. Because the circuit contains no closed loops, no current flows anywhere in the circuit. With no voltage drop across R1 , it follows that

+

+ _υC

C

L υL

_

t=0

 • The voltage υC across a capacitor cannot change instantaneously, and neither can the current iL through an inductor. • In circuits containing dc sources, the steady state condition of the circuit (after all transients have died out) is such that no currents flow through capacitors and no voltages exist across inductors, allowing us to represent capacitors as open circuits and inductors as short circuits under steady state conditions. 

RLC CIRCUITS

1

2 (a) Circuit R2 iL(0−) = 0

R1 Vs

C

+ + __

1

+ − _υC(0 ) = Vs

L

2

(b) At t = 0−, C acts like an open circuit and L like a short circuit R2 iL

iC C

+ _ Vs

L

iL(0) = 0

2 (c) At t = 0, C acts like a voltage source and L like a current source with zero current Figure 6-2: Circuit of Example 6-1.

υC (0− ) = Vs . Also, −

iL (0 ) = 0. Time-continuity of υC and iL mandates that after moving the switch to terminal 2: υC (0) = υC (0− ) = Vs , iL (0) = iL (0− ) = 0.

(b) The circuit in Fig. 6-2(c) depicts the state of the circuit at t = 0 (after moving the switch). The capacitor behaves like a dc voltage source of magnitude Vs , and the inductor behaves like a dc current source with zero current, which is equivalent to an open circuit. Even though in general there is no requirement disallowing a sudden change in iC , in this case iC = iL and iL (0) = 0. Consequently, iC (0) = 0.

“book” — 2015/5/4 — 7:16 — page 333 — #4

6-1

INITIAL AND FINAL CONDITIONS

333

I0 u(t) R2

iL

R1

R3

+ + __

C

(a)

+ _υC

R1 V0

+ + __

R1 V0

(c)

_

I0

+

_

2

R2

R3

C

i2

iL( ) L

iR3(0)

iL(0) = 2 A

+ + __

iC(0)

8

υL(0)

+ _ υC(0) = 12 V

R1

V0

At t = 0

R3

i1

+ + __

t=

(d)

+

υC( ) 8

i1(0)

C υC(0−) = 12 V

At t = 0−

I0

1

+

R3

(b)

R2

iC(0−) = 0

_

8

V0

iL(0−) = 2 A

R2

L

Figure 6-3: Circuit for Example 6-2.

With no voltage drop across R2 , the voltage across the inductor is

C = 8 mF. Determine: (a) υC (0) and iL (0), (b) iC (0) and υL (0), and (c) υC (∞) and iL (∞).

υL (0) = υC (0) = Vs .

Solution: (a) To find initial values of υC and iL at t = 0, we have to determine their values at t = 0− , and then invoke the requirement that neither the voltage across a capacitor nor the current through an inductor can change in zero time. The state of the circuit at t = 0− is shown in Fig. 6-3(b), wherein the inductor has been replaced with a short circuit, the capacitor replaced with an open circuit, and the current source is absent altogether. Since iC (0− ) = 0,

(c) The analysis for υC and iL as t → ∞ is totally straightforward; with no active sources remaining in the part of the circuit that contains L and C, all of the energy that may have been stored in L and C will have dissipated completely by t = ∞, rendering the circuit inactive. Hence, υC (∞) = 0,

iL (∞) = 0.

iL (0− ) = Example 6-2: Initial and Final Conditions

The circuit in Fig. 6-3(a) contains a dc voltage source and a step-function current source. The element values are V0 = 24 V, I0 = 4 A, R1 = 2 , R2 = 4 , R3 = 6 , L = 0.2 H, and

V0 = 2 A, R1 + R 2 + R 3

and υC (0− ) = iL (0− ) R3 = 12 V. Hence, iL (0) = iL (0− ) = 2 A,

“book” — 2015/5/4 — 7:16 — page 334 — #5

334

CHAPTER 6

and υC (0) = υC (0− ) = 12 V.

RLC CIRCUITS

Concept Question 6-3: What role do initial and final

values play in the solution of a circuit? (See

)

(b) At t = 0, the state of the circuit is as shown in Fig. 6-3(c). Since υR3 (0) = υC (0) = 12 V, it follows that 12 = 2 A. 6 We did this because we need iC (0). Application of KCL at node 2 leads to

Exercise 6-1: For the circuit in Fig. E6.1, determine υC (0), iL (0), υL (0), iC (0), υC (∞), and iL (∞).

iL

υL



iR3 (0) =

iC

L

+ 10 V _

t=0

υC

C



iC (0) = I0 + iL (0) − iR3 (0) = 4 + 2 − 2 = 4 A. Figure E6.1

Next, we need to determine υL (0). At node 1, i1 (0) = I0 + iL (0) = 4 + 2 = 6 A.

Answer: υC(0) = 6 V,

iL(0) = 1 A, υL(0) = −6 V, iC(0) = 0, υC(∞) = 0, iL(∞) = 0. (See C3 )

By applying KVL around the lower left loop, we find that υL (0) = −8 V. (c) The state of the circuit at t = ∞ shown in Fig. 6-3(d) resembles that at t = 0− , except that now we also have the current source I0 . The mesh equation for loop 1 is

Exercise 6-2: For the circuit in Fig. E6.2, determine υC (0), iL (0), υL (0), iC (0), υC (∞), and iL (∞).

iL

υC

C



L

iC

−V0 + R1 i1 + R2 (i1 − i2 ) + R3 i1 = 0, and for loop 2,

υL

t=0 + _ 12 V



i2 = I0 = 4 A. Solving for i1 gives

Figure E6.2

i1 = 3.33 A, which leads to

Answer: υC(0) = 0, iL(0) = 0, υL(0) = −12 V,

iC(0) = 0, υC(∞) = 4 V, iL(∞) = −2 A. (See

)

iL (∞) = i1 − I0 = 3.33 − 4 = −0.67 A and

6-2

Introducing the Series RLC Circuit

6-2.1

Charging-Up Mode

υC (∞) = i1 R3 = 3.33 × 6 = 20 V. Concept Question 6-1: Determination of initial circuit

conditions after a sudden change relies on two fundamental properties of capacitors and inductors. What are they? (See )

Concept Question 6-2: Under dc steady state conditions, does a capacitor resemble an open circuit or a short circuit? What does an inductor resemble? (See )

The circuit in part (a) of Fig. 6-4 depicts a scenario in which a series RLC circuit with no stored energy is connected to a dc voltage source Vs at t = 0. After closing the switch, charge supplied by the source starts to flow to the (+) voltage terminal of the capacitor, and continues to do so until the capacitor reaches the maximum voltage possible, namely Vs . Hence, our expectation is that υC (t) will start at zero at t = 0 and then build up to reach Vs as t → ∞. The specific path it takes,

“book” — 2015/5/4 — 7:16 — page 335 — #6

6-2

INTRODUCING THE SERIES RLC CIRCUIT

335

υC(t) L

R Vs

+ _

Critically damped (α = ω0)

iC

t=0 24 V

C

Underdamped (α < ω0)

Vs = 24 υC Overdamped (α > ω0)

υC(0−) = 0 0

0.05

(a) Charging up C

Vs

+ _

2

24 20

L

R

0.15

0.2

t (s)

(b) Responses υC(t)

1

0.1

Underdamped (α < ω0) Critically damped (α = ω0) Overdamped (α > ω0)

t=0 C

υC

10 0

υC(0−)

= 24 V

0

0.05

0.1

0.15

0.2

t (s)

−10 (c) Discharging C

(d) Responses

Figure 6-4: Illustrating the charge-up and discharge responses of a series RLC circuit with Vs = 24 V. In all cases R = 12  and L = 0.3 H, which specifies α = R/2L = 20 Np/s. When C = 0.01 F, the response is overdamped, when C = 8.33 mF, the response is critically damped, and when C = 0.72 mF, the response is underdamped. however, depends on the relative magnitudes of two important parameters. These are:

R damping coefficient α = 2L resonant frequency ω0 = √

(Np/s), 1

LC

(rad/s).

(6.1a)

Figure 6-4(b) displays three different response curves for υC (t), labeled as follows: Overdamped response Critically damped response Underdamped response

α > ω0 , α = ω0 , α < ω0 .

(6.1b)

(series RLC)

The parameter α is measured in nepers/second (Np/s) and ω0 is an angular frequency, measured in radians per second (rad/s). The magnitudes of the two parameters are specified by the values chosen for R, L, and C.

The critically damped response represents the fastest smooth path for υC (t) between its initial and final values. In comparison, the overdamped response is slower than the underdamped response, which starts out faster but exhibits an oscillatory (ringing) behavior. The mathematical solutions for all three cases are presented in detail in forthcoming sections. The intent is to provide an overview of how υC (t) varies with time under these various scenarios.

“book” — 2015/5/4 — 7:16 — page 336 — #7

336

CHAPTER 6

L

R + _

C

υC(0−) = 12 V

+ _

t=0 24 V

C

i

i

L R

+ _ 24 V

R

L

C C

(b) After t = 0

+ _

υC(0−) = 36 V

(a) At t = 0−

(a) At t = 0−

i=0

L

R

t=0 24 V

RLC CIRCUITS

R

L

24 V

υC increases from 12 V to reach 24 V after a long time

+ _

24 V

(b) After t = 0 R

C

υC(∞) = 24 V

υC decreases from 36 V to reach 24 V after a long time

+ _

L

24 V

i=0 C

υC(∞) = 24 V

(c) Long after closing the switch

(c) Long after closing the switch

Figure 6-5: Connecting a series RLC circuit with a charged-up capacitor to a source with higher voltage.

Figure 6-6: Connecting a series RLC circuit with a charged-up capacitor to a source with lower voltage.

6-2.2

Discharging Mode

If instead of starting out with an uncharged RLC circuit, we were to start with a fully charged capacitor, as depicted by the circuit in Fig. 6-4(c), and then discharge it by moving the SPDT switch from terminal 1 to terminal 2, the voltage υC (t) across the capacitor will decay from its initial value, Vs , to a final value of zero volts. The specific path between Vs and zero again depends on the value of α relative to that of ω0 , as shown in Fig. 6-4(d). In fact, the three responses of the discharging RLC circuit are essentially mirror images of those for the charging-up circuit; the initial and final conditions of the circuit in Fig. 6-4(a) are the converse of those for the circuit in Fig. 6-4(c). The capacitor voltage of the changing-up circuit starts at zero and concludes at 24 V, in contrast to the discharging circuit that starts at 24 V and concludes at zero. Now let us consider an RLC circuit in which the capacitor has 12 V across it (due to some previous charging-up action), and then a switch is closed to connect the RLC segment to a

source with Vs = 24 V, as shown in Fig. 6-5(a).After closing the switch (Fig. 6-5(b)), the situation is such that Vs = 24V exceeds the initial voltage of 12 V across the capacitor. Consequently, charge will flow to the capacitor to build up its voltage, and will continue to do so until the capacitor reaches the maximum possible voltage, namely Vs = 24 V. When it reaches that state, the current goes to zero (Fig. 6-5(c)). The scenario in Fig. 6-6 depicts a similar circuit, but one that starts with a capacitor whose initial voltage υC (0− ) is 36 V, which is higher than that of Vs = 24 V. In this case, the capacitor will start to discharge after closing the switch and then continue to discharge until it reaches 24 V. Thus, in both circuit scenarios, the capacitor will charge up or discharge down so as to equalize its voltage to that of the source, Vs . Recall that a short circuit is equivalent to a voltage source with Vs = 0. Hence, if we connect an RLC circuit with a charged-up capacitor to a short circuit, the capacitor will discharge down until it reaches a final voltage of zero, the same as the scenario depicted in Fig. 6-4(c).

“book” — 2015/5/4 — 7:16 — page 337 — #8

TECHNOLOGY BRIEF 15: MICROMECHANICAL SENSORS AND ACTUATORS

Technology Brief 15 Micromechanical Sensors and Actuators Energy is stored in many different forms in the world around us. The conversion of energy from one form to another is called transduction. Each of our five senses, for example, transduces a specific form of energy into electrochemical signals: tactile transducers on the skin convert mechanical and thermal energy; the eye converts electromagnetic energy; smell and taste receptors convert chemical energy; and our ears convert the mechanical energy of pressure waves. Any device, whether natural or man-made, that converts energy signals from one form to another is a transducer. Most modern man-made systems are designed to manipulate signals (i.e., information) using electrical energy. Computation, communication, and storage of information are examples of functions performed mostly with electrical circuits. Most systems also perform a fourth class of signal manipulation: the transduction of energy from the environment into electrical signals that circuits can use in support of their intended application. If a transducer converts external signals into electrical signals, it is called a sensor. The charge-coupled device (CCD) chip on your camera is a sensor that converts electromagnetic energy (light) into electrical signals that can be processed, stored, and communicated by your camera circuits. Some transducers perform the reverse function, namely to convert a circuit’s electrical signal into an environmental excitation. Such a transducer then is called an actuator. The components that deploy the airbag in your car are actuators: given the right signal from the car’s microcontroller, the actuators convert electrical energy into mechanical energy and the airbag is released and inflated.

Microelectromechanical Systems (MEMS) Micro- and nanofabrication technology have begun to revolutionize many aspects of sensor and actuator design. Humans increasingly are able to embed transducers at very fine scales into their environment. This is leading to big changes, as our computational elements are becoming increasingly aware of their environment. Shipping containers that track their own acceleration profiles, laptops that scan fingerprints for routine login, cars that detect collisions, and even office suites that modulate energy consumption based on human activity are all examples of this transduction revolution. In this

337

technology brief, we will focus on a specific type of microscale transducers that lend themselves to direct integration with silicon ICs. Collectively, devices of this type are called microelectromechanical systems (MEMS) or microsystems technologies (MST); the two names are used interchangeably.

A Capacitive Sensor: The MEMS Accelerometer According to Eq. (5.21), the capacitance C of a parallel plate capacitor varies directly with A, the effective area of overlap between its two conducting plates, and inversely with d, the spacing between the plates. By capitalizing on these two attributes, capacitors can be made into motion sensors that can measure velocity and acceleration along x, y, and z. Figure TF15-1 illustrates two mechanisms for translating motion into a change of capacitance. The first generally is called the gap-closing mode, while the second one is called the overlap mode. In the gap-closing mode, A remains constant, but if a vertical force is applied onto the upper plate, causing it to be displaced from its nominal position at height d above the lower plate to a new position (d − z), then the value of capacitance Cz will change in accordance with the expression given in Fig. TF15-1(a). The sensitivity of Cz to the vertical displacement is given by dCz /dz. The overlap mode (Fig. TF15-1(b)) is used to measure horizontal motion. If a horizontal force causes one of the plates to shift by a distance y from its nominal position (where nominal position corresponds to a 100 percent overlap), the decrease in effective overlap area will lead to a corresponding change in the magnitude of capacitance Cy . In this case, d remains constant, but the width of the overlapped areas changes from w to (w − y). The expression for Cy given in Fig. TF15-1(b) is reasonably accurate (even though it ignores the effects of the fringing electric field between the edges of the two plates) so long as y  w. To measure and amplify changes in capacitance, the capacitor can be integrated into an appropriate op-amp circuit whose output voltage is proportional to C. As we shall see shortly, a combination of three capacitors, one to sense vertical motion and two to measure horizontal motion along orthogonal axes, can provide complete information on both the velocity and acceleration vectors associated with the applied force. The capacitor configurations shown in Fig. TF15-1 illustrate the basic concept of how a capacitor is used to measure motion, although more complex capacitor

“book” — 2015/5/4 — 7:16 — page 338 — #9

338

TECHNOLOGY BRIEF 15: MICROMECHANICAL SENSORS AND ACTUATORS

Gap-Closing Mode

Overlap Mode y

Nominal position (no force)

Metal plates

F

z d

l

w Capacitance: Sensitivity:

εwl d−z εwl dCz = dz (d − z)2

Cz =

F Metal d plates l

w Capacitance: Sensitivity:

ε(w − y)l d dCy εl =− dy d

Cy =

Figure TF15-1: Basic capacitive measurement modes. For (b), the expressions hold only for small displacements such that y  w.

geometries also are possible, particularly for sensing angular motion. To convert the capacitor-accelerometer concept into a practical sensor—such as the automobile accelerometer that controls the release of the airbag—let us consider the arrangement shown in Fig. TF15-2(b). The lower plate is fixed to the body of the vehicle, and the upper plate sits on a plane at a height d above it. The upper plate is attached to the body of the vehicle through a spring with a spring constant k. When no horizontal force is acting on the upper plate, its position is such that it provides a 100 percent overlap with the lower plate, in which case the capacitance will be a maximum at Cy = εW/d. If the vehicle accelerates in the y-direction with acceleration ay , the acceleration force Facc will generate an opposing spring force Fsp of equal magnitude. Equating the two forces leads to an expression relating the displacement y to the acceleration ay , as shown in the figure. Furthermore, the capacitance Cy is directly proportional to the overlap area (w − y) and therefore is proportional to the acceleration ay . Thus, by measuring Cy , the accelerometer determines the value of ay . A similar overlap-mode capacitor attached to the vehicle along the x-direction can be used to measure ax . Through a similar analysis for the gap-closing mode capacitor shown in Fig. TF15-2(a), we can arrive at a functional relationship that can be used to determine the vertical acceleration az by measuring capacitance Cz . For example, if we designate the time when the ignition starts the engine as t = 0, we then can set the initial

conditions on both the velocity u of the vehicle and its acceleration a as zero at t = 0. That is, u(0) = a(0) = 0. The capacitor accelerometers measure continuous-time waveforms ax (t), ay (t), and az (t). Each waveform then can be used by an op-amp integrator circuit to calculate the corresponding velocity waveform. For ux , for example, t ux (t) =

ax (t) dt, 0

and similar expressions apply to uy and uz .

Commercial MEMS Accelerometers Figure TF15-3 shows the Analog Devices ADXL202 accelerometer which uses the gap-closing mode to detect accelerations on a tiny micromechanical capacitor structure that works on the same principle described above, although slightly more complicated geometrically. Commercial accelerometers, such as this one, make use of negative feedback to prevent the plates from physically moving. When an acceleration force attempts to move the plate, an electric negative-feedback circuit applies a voltage across the plates to generate an electrical force between the plates that counteracts the acceleration force exactly, thereby preventing any motion by the plate. The magnitude of the applied voltage becomes a measure of the acceleration force that the capacitor plate is subjected to. Because of their small size and low power consumption, chip-based microfabricated silicon accelerometers

“book” — 2015/5/4 — 7:16 — page 339 — #10

TECHNOLOGY BRIEF 15: MICROMECHANICAL SENSORS AND ACTUATORS

Fsp

339

Spring constant k Fsp

Facc w

z

Mass m

d

Facc

y Spring

d

Fsp = Facc

Facc = Fsp

kz = maz

may = ky may y= k

z=

maz k

εwl Cz = maz d− k

(

)

(a) The ADXL202 accelerometer employs many gap-closing capacitor sensors to detect acceleration. (Courtesy Analog Devices.)

Cy =

εl(w − y) = d

(

εl w −

may k

)

d

(b) A silicon sensor that uses overlap mode fingers. The white arrow shows the direction of motion of the moving mass and its fingers in relation to the fixed anchors. Note that the moving fingers move into and out of the fixed fingers on either side of the mass during motion. (Courtesy of the Adriatic Research Institute.)

Figure TF15-2: Adding a spring to a movable plate capacitor makes an accelerometer. are used in most modern cars to activate the release mechanism of airbags.They also are used heavily in many toy applications to detect position, velocity and acceleration. The Nintendo Wii, for example, uses accelerometers in each remote to detect orientation and acceleration. Incidentally, a condenser microphone operates much like the device shown in Fig. TF15-2(a): as air pressure waves

(sound) hit the spring-mounted plate, it moves and the change in capacitance can be read and recorded.

A Capacitive Actuator: MEMS Electrostatic Resonators Not surprisingly, we can drive the devices discussed previously in reverse to obtain actuators. Consider again

“book” — 2015/5/4 — 7:16 — page 340 — #11

340

TECHNOLOGY BRIEF 15: MICROMECHANICAL SENSORS AND ACTUATORS

FigureTF15-3: The complete ADXL202 accelerometer chip.The center region holds the micromechanical sensor; the majority of the chip space is used for the electronic circuits that measure the capacitance change, provide feedback, convert the measurement into a digital signal, and perform self-tests. (Courtesy of Analog Devices.)

the configuration in Fig. TF15-2(a). If the device is not experiencing any external forces and we apply a voltage V across the two plates, an attractive force F will develop between the plates. This is because charges of opposite polarity on the two plates give rise to an electrostatic force between them. This, in fact, is true for all capacitors. In the case of our actuator, however, we replace the normally stiff, dielectric material with air (since air is itself a dielectric) and attach it to a spring as before. With this modification, an applied potential generates an electrostatic force that moves the plates. This basic idea can be applied to a variety of applications. A classic application is the digital light projector (DLP) system that drives most digital projectors

used today. In the DLP, hundreds of thousands of capacitor actuators are arranged in a 2-D array on a chip, with each actuator corresponding to a pixel on an image displayed by the projector. One capacitive plate of each pixel actuator (which is mirror smooth and can reflect light exceedingly well) is connected to the chip via a spring. In order to brighten or darken a pixel, a voltage is applied between the plates, causing the mirror to move into or out of the path of the projected light. These same devices have been used for many other applications, including microfluidic valves and tiny force sensors used to measure forces as small as a zeptonewton (1 zeptonewton = 10−21 newtons).

“book” — 2015/5/4 — 7:16 — page 341 — #12

6-3

SERIES RLC OVERDAMPED RESPONSE (α > ω0 )

R Vs

+ _

L

t=0 24 V

and rearranging terms, Eq. (6.2) becomes

iC C

341

R dυC Vs 1 d 2 υC + + υC = . (6.4) 2 dt L dt LC LC For convenience, we rewrite Eq. (6.4) in the abbreviated form

υC

υC + aυC + bυC = c,

Figure 6-7: Series RLC circuit connected to a source Vs at

t = 0. In general, the capacitor may have had an initial charge on it at t = 0− , with a corresponding initial voltage υC (0− ).

6-3 Series RLC Overdamped Response (α > ω0 ) A key takeaway lesson from the qualitative description given in the preceding section is that after closing the switch in a series RLC circuit, the voltage across the capacitor will charge up or discharge down to equalize to the voltage across the source. In this section, we derive the differential equation for the series RLC circuit in Fig. 6-7 and then solve it to obtain an expression for υC (t) for t ≥ 0, with t = 0 designated as the time immediately after the switch is closed. As noted in the preceding section, the nature of the solution for υC (t) depends on how the magnitude of the damping coefficient α compares with that of the resonant frequency ω0 . The values of the two parameters are dictated by the values of R, L, and C, per the expressions in Eq. (6.1). In the present section, we consider the case corresponding to α > ω0 , which is called the overdamped response. The other two cases are treated in follow-up sections.

6-3.1

For the circuit in Fig. 6-7, the KVL loop equation for t ≥ 0 (after closing the switch) is RiC + L

diC + υC = Vs dt

(for t ≥ 0),

(6.2)

where iC and υC are the current through and voltage across the capacitor. The capacitor may or may not have had charge on it. If it had, we denote the value of the initial voltage across it υC (0), which is the same as υC (0− ), the voltage across it before closing the switch (since the voltage across a capacitor cannot change instantaneously). By incorporating the relation dυC iC = C , dt

where

R 1 Vs , b= , c= . (6.6) L LC LC The second-order differential equation given by Eq. (6.5) is specific to the capacitor voltage of the series RLC circuit of Fig. 6-7, but the form of the equation is equally applicable to any current or voltage in any second-order circuit (although the values of the constants a, b, and c are different for different circuits). The same is true for the general form of the solution of the differential equation. a=

6-3.2

Solution of Differential Equation

The general solution of the second-order differential equation given by Eq. (6.5) consists of two components: υC (t) = υtr (t) + υss (t),

(source-free).

(6.8)

The steady-state solution υss (t) is related to the forcing function on the right-hand side of Eq. (6.5), and its functional form is similar to that of the forcing function. Since in the present case, the forcing function c is simply a constant, so is υss (t). That is, υss (t) is a non–time-varying constant υss that will be determined later from initial and final conditions. Moreover, as we will see shortly, the transient component υtr (t) always goes to zero as t → ∞ (that’s why it is called transient). Hence, as t → ∞, Eq. (6.7) reduces to υC (∞) = υss ,

(6.9)

in which case Eq. (6.7) can be rewritten as υC (t) = υtr (t) + υC (∞).

(6.3)

(6.7)

where υtr (t) is the transient (also called homogeneous solution of Eq. (6.5) or the natural response of the RLC circuit) and υss (t) is the steady-state solution (also called particular solution). The transient solution is the solution of Eq. (6.5) under source-free conditions; i.e., with Vs = 0, which means that c = Vs /LC also is zero. Thus υtr (t) is the solution of υtr + aυtr + bυtr = 0

Differential Equation

(6.5)

Our remaining task is to determine υtr (t).

(6.10)

“book” — 2015/5/4 — 7:16 — page 342 — #13

342

CHAPTER 6

When differentiated, the exponential function est replicates itself (within a multiplying factor), so it is often offered as a candidate solution when solving homogeneous differential equations. Thus, we assume that υtr (t) = Aest ,

R a = 2L 2 1 =b ω0 = √ LC α=

(6.11)

(rad/s),

the expressions given by Eq. (6.14) become

s 2 Aest + asAest + bAest = 0,

s2 = −α −

(6.12)

 

(6.17a)

(Np/s),

where A and s are constants to be determined later. To ascertain that Eq. (6.11) is indeed a viable solution of Eq. (6.8), we insert the proposed expression for υtr (t) and its first and second derivatives in Eq. (6.8). The result is

s1 = −α +

RLC CIRCUITS

(6.17b)

α 2 − ω02 ,

(6.18a)

α 2 − ω02 ,

(6.18b)

which simplifies to s 2 + as + b = 0.

(6.13)

Hence, the proposed solution given by Eq. (6.11) is indeed an acceptable solution so long as Eq. (6.13) is satisfied. The quadratic equation given by Eq. (6.13) is known as the characteristic equation of the differential equation. It has two roots:   a 2 −b , 2   a 2 a s2 = − − −b . 2 2 a s1 = − + 2

(6.14a)

for t ≥ 0,

Invoking Initial Conditions

To determine the values of constants A1 and A2 in Eq. (6.16), we need to invoke initial conditions, which means that we need to use information available to us about the values of υC and its time derivative υC , both at t = 0. Since iC (t) = C

(6.16)

The exponential coefficients s1 and s2 are given by Eq. (6.14) in terms of constants a and b, both of which are defined in Eq. (6.6). By reintroducing the damping coefficient α and resonant frequency ω0 , which we defined earlier in Eq. (6.1), as

dυC = C υ  (t), dt

(6.19)

the second requirement is equivalent to needing to know iC (0). At t = 0, Eq. (6.16) simplifies to υC (0) = A1 + A2 + υC (∞), and iC (0) = C

(6.20)

  dυC  = C(s1 A1 es1 t + s2 A2 es2 t )t=0  dt t=0 = C(s1 A1 + s2 A2 ).

(6.15)

where constants A1 and A2 are to be determined shortly. Inserting Eq. (6.15) into Eq. (6.10) leads to υC (t) = A1 es1 t + A2 es2 t + υC (∞).

6-3.3

(6.14b)

Since the values of a and b are governed by the values of only the passive components in the circuit, so are the values of s1 and s2 . Strictly speaking, the unit of s1 and s2 is 1/second, but it is customary to add the dimensionless neper to the units of quantities that appear in exponential functions. Hence, s1 and s2 are measured in nepers/second (Np/s). The existence of two distinct roots implies that Eq. (6.8) has two viable solutions, one in terms of es1 t and another in terms of es2 t . Hence, we should generalize the form of our solution to υtr (t) = A1 es1 t + A2 es2 t

The solution in the present section pertains to the overdamped case corresponding to α > ω0 . Under this condition, both s1 and s2 are real, negative numbers. Consequently, as t → ∞, the first two terms in Eq. (6.16) go to zero, just as we asserted earlier.

(6.21)

Simultaneous solution of Eqs. (6.20) and (6.21) for A1 and A2 gives A1 = A2 =

1 C iC (0) − s2 [υC (0) − υC (∞)]

s1 − s 2

1 C iC (0) − s1 [υC (0) − υC (∞)]

s2 − s 1

,

(6.22a)

.

(6.22b)

This concludes the general solution for the overdamped response. A summary of relevant expressions is available in Table 6-1.

“book” — 2015/5/4 — 7:16 — page 343 — #14

6-3

SERIES RLC OVERDAMPED RESPONSE (α > ω0 )

343

Table 6-1: Step response of RLC circuits for t ≥ 0. Series RLC

R

Input: dc circuit with switch action @t=0

Parallel RLC

L C

υC

Input: dc circuit with switch action @t=0

Total Response

R

L

C

Total Response

Overdamped (α > ω0 )

Overdamped (α > ω0 )

υC (t) = A1 es1 t + A2 es2 t + υC (∞)

iL (t) = A1 es1 t + A2 es2 t + iL (∞)

1 i (0) − s [υ (0) − υ (∞)] C 2 C C A1 = C s1 − s 2 1 C iC (0) − s1 [υC (0) − υC (∞)] A2 = s2 − s 1

Critically Damped (α = ω0 )

1 υ (0) − s [i (0) − i (∞)] L 2 L L A1 = L s1 − s 2  1 υ (0) − s [i (0) − i (∞)] L 1 L L L A2 = s2 − s 1

Critically Damped (α = ω0 )

υC (t) = (B1 + B2 t)e−αt + υC (∞)

iL (t) = (B1 + B2 t)e−αt + iL (∞)

B1 = υC (0) − υC (∞)

B1 = iL (0) − iL (∞)

B2 = C1 iC (0) + α[υC (0) − υC (∞)]

1 υ (0) + α[i (0) − i (∞)] B2 = L L L L

Underdamped (α < ω0 )

Underdamped (α < ω0 )

υC (t) = e−αt (D1 cos ωd t + D2 sin ωd t) + υC (∞)

α=

iL

iL (t) = e−αt (D1 cos ωd t + D2 sin ωd t) + iL (∞)

D1 = υC (0) − υC (∞)

D1 = iL (0) − iL (∞)

1 i (0) + α[υ (0) − υ (∞)] C C C D2 = C ωd

1 υ (0) + α[i (0) − i (∞)] L L L D2 = L ωd

Auxiliary Relations

⎧ R ⎪ ⎪ ⎨ 2L ⎪ ⎪ ⎩ 1 2RC

s1 = −α +

Series RLC Parallel RLC 

α 2 − ω02

Example 6-3: Charging Up Capacitor with No Prior Charge

Given that in the circuit of Fig. 6-8(a), Vs = 16 V, R = 64 , L = 0.8 H, and C = 2 mF, determine υC (t) and iC (t) for t ≥ 0. The capacitor had no charge prior to t = 0.

1 ω0 = √ LC  ωd = ω02 − α 2  s2 = −α − α 2 − ω02

Solution: We begin by establishing the damping condition of the circuit. From the definitions for α and ω0 given by Eq. (6.17), we have R 64 = = 40 Np/s, 2L 2 × 0.8 1 1 =√ ω0 = √ = 25 rad/s. LC 0.8 × 2 × 10−3 α=

“book” — 2015/5/4 — 7:16 — page 344 — #15

344

CHAPTER 6

L

R + _

Vs

RLC CIRCUITS

As t → ∞, the circuit reaches steady state and the capacitor becomes like an open circuit, allowing no current to flow through the circuit. Consequently,

i

t=0 υC

C

υC (∞) = Vs = 16 V. At t = 0− , the capacitor was uncharged. Hence,

(a)

υC (0) = υC (0− ) = 0.

υC (V)

12

Prior to t = 0, there was no current in the circuit, and since the current through L (which is also the current through C) cannot change instantaneously, it follows that

8

iC (0) = iL (0) = iL (0− ) = 0.

16

Capacitor voltage

From Eq. (6.22), A1 and A2 are given by

4

0

0.1

0.2

0.3

0.4

0.5

t (s)

(b) iC (A) 0.2 0.15 Current

0.1

1 C iC (0) − s2 [υC (0) − υC (∞)]

s1 − s 2 0 + 71.2(0 − 16) = −18.25 V, = −8.8 + 71.2  1 C iC (0) − s1 [υC (0) − υC (∞)] A2 = − s1 − s 2   0 + 8.8(0 − 16) =− = 2.25 V. −8.8 + 71.2 The total response υC (t) is then given by

0.05 0 0

A1 =

t (s) 0.1

0.2

0.3

0.4

Figure 6-8: Example 6-3: (a) circuit, (b) υC (t), and (c) iC (t). Hence, α > ω0 , which means that the circuit will exhibit an overdamped response after the switch is closed. The applicable expression for υC (t) is given by Eq. (6.16),

From Eq. (6.18), s1 = −α +



α 2 − ω02  = −40 + 402 − 252 = −8.8 Np/s,  s2 = −α − α 2 − ω02 = −71.2 Np/s.

(for t ≥ 0),

0.5

(c)

υC (t) = [A1 es1 t + A2 es2 t + υC (∞)].

υC (t) = [−18.25e−8.8t + 2.25e−71.2t + 16] V

and the associated current is iC (t) = C

dυC dt

= 2 × 10−3 [18.25 × 8.8e−8.8t − 2.25 × 71.2e−71.2t ] = 0.32(e−8.8t − e−71.2t ) A

(for t ≥ 0).

The waveforms of υC (t) and iC (t) are displayed in Figs. 6-8(b) and (c), respectively. Example 6-4: RLC Circuit with a Current Source

Determine υC (t) in the circuit of Fig. 6-9(a), given that Is = 2 A, Rs = 10 , R1 = 1.81 , R2 = 0.2 , L = 5 mH,

“book” — 2015/5/4 — 7:16 — page 345 — #16

6-3

SERIES RLC OVERDAMPED RESPONSE (α > ω0 )

R1

2

345

R1

1 iL(0−)

1.81 Ω t = 0

=0

L υC Rs

C 8V

10 Ω

2

1.81 Ω

10 Ω

L

Is = 2 A

R2 0.2 Ω

Rs

1

υC(0) = 20 V

C

+ _

RsIs = 20 V

R2 0.2 Ω

8V

(b) At t = 0− (after current-to-voltage transformation)

(a) Original circuit

υC (V) 25 R1 iL L 8V

2

20

1.81 Ω

15

υC

C

R2 0.2 Ω υC(0−) = 20 V

10 8 5 0

0

5

10

15

(c) After t = 0

20

25

30

t (ms)

(d) υC(t) Figure 6-9: Circuit for Example 6-4.

and C = 5 mF. Assume that the circuit had been in the condition shown in Fig. 6-9(a) for a long time prior to t = 0. Solution: At t = 0− : Figure 6-9(b) depicts the state of the circuit at t = 0− , but after making a current source to voltage source transformation. The replacement voltage source is 20 V. Since the circuit had been in steady state for a long time, the capacitor behaves like an open circuit with

of R = R1 + R2 = 1.81 + 0.2 = 2.01 , L = 5 mH, and C = 5 mF, all connected in series with an 8 V source (Fig. 6-9(c)). The current through C is the same as the current through L, and since the current through an inductor cannot charge instantaneously, it follows that iC (0) = iL (0) = iL (0− ) = 0. For the capacitor, υC (0) = υC (0− ) = 20 V.

υC (0− ) = 20 V.

Also, as t approaches ∞, υC (t) approaches the voltage of the 8 V source. Hence,

We also note that in the left-hand part of the circuit, no current can flow, mandating that

υC (∞) = 8 V.

iL (0− ) = 0. At t ≥ 0: After moving the switch to terminal 2, the capacitor becomes part of a new circuit composed of a combination

The parameters α and ω0 are given by 2.01 R1 + R2 = 201 Np/s, = 2L 2 × 5 × 10−3 1 1 =√ ω0 = √ = 200 rad/s. −3 LC 5 × 10 × 5 × 10−3 α=

“book” — 2015/5/4 — 7:16 — page 346 — #17

346

CHAPTER 6

Since α > ω0 , the response is overdamped and given by Eq. (6.16),

RLC CIRCUITS

α = ω0 , and according to Eq. (6.18), s1 = s2 = −α.

υC (t) = A1 es1 t + A2 es2 t + υC (∞),

(6.24)

Repeated roots are problematic because Eq. (6.16) becomes

with s1 = −α +



υC (t) = A1 e−αt + A2 e−αt + υC (∞)

α 2 − ω02  = −201 + (201)2 − (200)2 = −181 Np/s,  s2 = −α − α 2 − ω02 = −221 Np/s, A1 =

= (A1 + A2 )e−αt + υC (∞) = (A3 )e−αt + υC (∞), (6.25) where A3 = A1 + A2 . A solution containing a single constant (A3 ) cannot simultaneously satisfy the initial conditions on both the voltage across the capacitor and the current through the inductor. For this critically damped case, we introduce two new constants, B1 and B2 , and we adopt the modified form

1 C iC (0) − s2 [υC (0) − υC (∞)]

s1 − s 2 0 + 221[20 − 8] = = 66.3, −181 + 221

A2 =

1 C iC (0) − s1 [υC (0) − υC (∞)]

s2 − s 1 0 + 181[20 − 8] = = −54.3. −221 + 181

υC (t) = B1 e−αt + B2 te−αt + υC (∞) = (B1 + B2 t)e−αt + υC (∞)

Inserting the values of s1 , s2 , A1 , A2 , and υC (∞) in Eq. (6.16) leads to υC (t) = (66.3e−181t − 54.3e−221t + 8) V

for t ≥ 0.

Figure 6-9(d) displays the time response of υC (t). Exercise 6-3: After interchanging the locations of L

and C in Fig. 6-9(a), repeat Example 6-4 to determine υC(t) across C. Answer: υ(t) = 9.8(e−221t − e−181t ) V. (See

C3

)

(for t ≥ 0)

(critically damped),

which contains a term with e−αt and a second term with (te−αt ). It is a relatively straightforward task to show that the expression given by Eq. (6.26) is indeed a valid solution of the differential equation given by Eq. (6.4). When doing so, however, we need to keep in mind that under the critically damped condition, R, L, and C are interrelated by Eq. (6.23), and υC (∞) = Vs . The constants B1 and B2 are governed by the initial conditions on υC and ic . Thus, at t = 0, Eq. (6.26) provides υC (0) = B1 + υC (∞),  dυC  iC (0) = C dt t=0

6-4 Series RLC Critically Damped Response (α = ω0 )

(6.26)

(6.27a)

 = C (−αB1 − αB2 t + B2 )e−αt t=0

 The critically damped response is the fastest response the circuit can exhibit, without oscillation, between initial and final conditions. 

= C(−αB1 + B2 ).

(6.27b)

Simultaneous solution of Eqs. (6.27a and b) leads to When  R=2

B1 = υC (0) − υC (∞), L C

(critically damped),

(6.23)

B2 =

1 iC (0) + α[υC (0) − υC (∞)]. C

(6.28a) (6.28b)

“book” — 2015/5/4 — 7:16 — page 347 — #18

SERIES RLC CRITICALLY DAMPED RESPONSE (α = ω0 )

6-4

L

R + _

Application of these initial and final conditions to Eq. (6.28) leads to

iC

t=0 Vs = 24 V

347

B1 = υC (0) − υC (∞) = −24 V,

υC

C

1 iC (0) + α[υC (0) − υC (∞)] C = 0 + 20[0 − 24] = −480.

B2 =

(a) Hence,

υC(t)

υC (t) = (B1 + B2 t)e−αt + υC (∞) = [−(24 + 480t)e−20t + 24] V,

24

for t ≥ 0.

The response is plotted in Fig. 6-10(b).

Critically damped (α = ω0)

Exercise 6-4: The switch in Fig. E6.4 is moved to position 2 after it had been in position 1 for a long time. Determine: (a) υC (0) and iC (0), and (b) iC (t) for t ≥ 0.

0

0.05

0.1

0.15

0.2

2

20 Ω

t (s)

1

t=0

(b) υC(t) 1H

iC Figure E6.4

Example 6-5: Critically Damped Response

Solution: The parameters α and ω0 are given by 12 R = = 20 Np/s, 2L 2 × 0.3 1 1 =√ ω0 = √ = 20 rad/s. LC 0.3 × 8.33 × 10−3 α=

Hence, because α = ω0 , the response is critically damped and given by Eq. (6.26) as υ(t) = (B1 + B2 t)e−20t + υC (∞).

Answer: (a) υC(0) = 40 V, iC(0) = 0.

(b) iC(t) = [−40te−10t ] A. (See

and

iC (0) = 0,

and the final condition on υC is υC (∞) = Vs = 24 V.

)

Exercise 6-5: The circuit in Fig. E6.5 is a replica of the circuit in Fig. E6.4, but with the capacitor and inductor interchanged in location. Determine: (a) iL (0) and υL (0), and (b) iL (t) for t ≥ 0.

20 Ω

2

1

10 Ω

t=0 10 mF

1H

υL iL

The initial conditions at t = 0 are υC (0) = 0

+ _ 40 V

υC

10 mF

Figure 6-10: Circuit response for Example 6-5.

Evaluate the response of the circuit in Fig. 6-10(a) for t ≥ 0, given that the capacitor had no charge prior to t = 0 and Vs = 24 V, R = 12 , L = 0.3 H, and C = 8.33 mF.

10 Ω

Figure E6.5 Answer: (a) iL(0) = 4 A,

υL(0) = −80 V. ) (b) iL(t) = [4(1 − 10t)e−10t ] A. (See

+ _ 40 V

“book” — 2015/5/4 — 7:16 — page 348 — #19

348

CHAPTER 6

The negative exponential e−αt signifies that υ(t) has a damped waveform with a time constant τ = 1/α, and the sine and cosine terms signify that υC (t) is oscillatory with an angular frequency ωd and a corresponding time period

6-5 Series RLC Underdamped Response (α < ω0 ) If α < ω0 , corresponding to

T =



L R 0. In terms of ωd , the expressions for the roots s1 and s2 given by Eq. (6.18) become 

RLC CIRCUITS

−αt j ωd t

e

+ A2 e

−αt −j ωd t

e

+ υC (∞).

(6.32)

The Euler identity e±j θ = cos θ ± j sin θ

The oscillatory behavior of the underdamped response is illustrated by Example 6-6. Example 6-6: Underdamped Response

Determine υC (t) for the circuit in Fig. 6-11, given that Vs = 24 V, R = 12 , L = 0.3 H, and C = 0.72 mF. The circuit had been in steady state prior to moving the switch at t = 0. Solution: For the specified values of R, L, and C,

(6.33) α=

allows us to expand Eq. (6.32) as follows:

12 R = = 20 Np/s 2L 2 × 0.3

and υC (t) = A1 e−αt (cos ωd t + j sin ωd t)

ω0 = √

+ A2 e−αt (cos ωd t − j sin ωd t) + υC (∞) = e−αt [(A1 + A2 ) cos ωd t + j (A1 − A2 ) sin ωd t] + υC (∞).

(6.34)

Next, by introducing a new pair of constants, D1 = A1 + A2 and D2 = j (A1 − A2 ), we have

(for t ≥ 0)

(underdamped). (6.35)

LC

=√

1 0.3 × 0.72 × 10−3

= 68 rad/s.

Since α < ω0 , the voltage response is underdamped and given by Eq. (6.35) as υC (t) = e−αt [D1 cos ωd t + D2 sin ωd t] + υC (∞), with ωd =

υC (t) = e−αt [D1 cos ωd t + D2 sin ωd t] + υC (∞)

1

  ω02 − α 2 = (68)2 − (20)2 = 65 rad/s.

Prior to t = 0, the circuit was in steady state, which means that the capacitor was fully charged at Vs = 24 V and acting like an open circuit. Hence, υC (0− ) = 24 V and iC (0− ) = 0.

“book” — 2015/5/4 — 7:16 — page 349 — #20

6-6

SUMMARY OF THE SERIES RLC CIRCUIT RESPONSE

R

1 Vs

+ _

Concept Question 6-4: What specific feature distin-

L

iC

t=0

2

349

C

υC

guishes the waveform of the underdamped response from those of the overdamped and critically damped responses? (See ) Concept Question 6-5: Why is ωd called the damping

frequency? (See

(a) υC (V)

)

Exercise 6-6: Repeat Example 6-4 after replacing the 8 V

source with a short circuit and changing the value of R1 to 1.7 .

24 20

Answer:

Underdamped

16

υ(t) = e−190t (20 cos 62.45t + 60.85 sin 62.45t) V.

12 8

(See

)

4 0 −4

0

0.05

0.1

0.15

0.2

t (s)

6-6

Summary of the Series RLC Circuit Response

6-6.1

Switch Action at t = 0

−8 −12 (b) Figure 6-11: Example 6-6 (a) circuit and (b) υC (t).

Since both υC across C and iL through L cannot change instantaneously, υC (0) = 24 V,

The left-hand column of Table 6-1 provides the general expressions for υC (t) for each of the three damping conditions associated with the series RLC circuit. The table also includes expressions for the constants in those expressions in terms of the initial and final values of υC and the initial value of iC . In all three cases, the starting point is to compute the values of α and ω0 , then their relative values determines the applicable damping condition.

6-6.2 −

iC (0) = iL (0) = iL (0 ) = 0. After t = 0, the closed RLC circuit will no longer have any active sources, allowing the capacitor to dissipate all its energy in the resistor. Hence, as t → ∞, υC (∞) = 0. Using these initial and final values in the appropriate expressions for D1 and D2 in Eq. (6.37) leads to D1 = 24 V, D2 = 7.4 V, and υC (t) = e−20t [24 cos 65t + 7.4 sin 65t] V,

for t ≥ 0.

Figure 6-11(b) shows a time plot of υC (t), which exhibits an exponential decay (due to e−20t ) in combination with the oscillatory behavior associated with the sine and cosine functions.

Switch Action at t = T0

If the sudden change in the circuit occurs at t = T0 , instead of at t = 0, the only changes that need to be made are: (1) t should be replaced with (t −T0 ) everywhere on the righthand side of all equations in Table 6-1. (2) υC (0) and iC (0) should be replaced with υC (T0 ) and iC (T0 ), respectively, in the expressions for the constants in Table 6-1.

Example 6-7: Rectangular-Pulse Excitation

The switch in the circuit of Fig. 6-12(a) was in position 1 for a long time before it was moved to position 2 at t = 0, and

“book” — 2015/5/4 — 7:16 — page 350 — #21

350

CHAPTER 6

2t=0 Vs

R

L

iC

υC (V)

t = 20 ms

+ _

RLC CIRCUITS

2.5

1

C

υC

2 1.5 1.08

(a)

Capacitor voltage

1 0.5 R

2

L

0

iC1

t (ms) 0

(d) + Vs u(t) _

C

20

40

60

80

100

Switch 2 1

υC1 iC (A)

0 < t < 20 ms

(b)

0.2 0.182 0.15 Current

0.1 R

L

0.05 t (ms)

0

1

(c)

iC2

C

υC2

After t = 20 ms

−0.05

(e)

0

20

40

60

80

100

Switch 2 1

Figure 6-12: Example 6-7 with Vs = 12 V, R = 40 , L = 0.8 H, and C = 2 mF.

then back to position 1 at t = 20 ms. If Vs = 12 V, R = 40 , L = 0.8 H, and C = 2 mF, determine the waveforms of υC (t) and i(t) for t ≥ 0. Solution: From Eq. (6.17), 40 R = = 25 Np/s, α= 2L 2 × 0.8 1 1 =√ ω0 = √ = 25 rad/s. LC 0.8 × 2 × 10−3 Since α = ω0 , the circuit will exhibit a critically damped response. We will divide the solution into two time segments.

Time Segment 1:

0 ≤ t ≤ 20 ms.

The general expression for the critically damped response of the series RLC circuit is given by Eq. (6.26) as υC1 (t) = (B1 + B2 t)e−αt + υ1 (∞).

(6.38)

Even though we know that the switch will be moved back to position 1 at t = 20 ms, when we evaluate the constants in Eq. (6.38) for Time Segment 1, we do so as if the state of the circuit shown in Fig. 6-12(b) is to remain the same until t = ∞. Since the circuit is “unaware” of the change that will be taking place at t = 20 ms, its reaction to the change at t = 0 presumes

“book” — 2015/5/4 — 7:16 — page 351 — #22

6-6

SUMMARY OF THE SERIES RLC CIRCUIT RESPONSE

that the new condition of the circuit will continue indefinitely. Hence, the voltage across the capacitor at t = ∞ would have been υC1 (∞) = Vs = 12 V.

(6.39)

At t = 0− , the RLC circuit contains no active sources, so both υ1 (0− ) and i1 (0− ) are zero. Moreover, since neither the voltage across C nor the current through L can change instantaneously, it follows that

351 where constants B3 and B4 are so labeled to avoid confusion with B1 and B2 of the earlier time segment. The associated current is iC2 (t) = C

dυC2 dt

d {[B3 + B4 (t − 0.02)]e−25(t−0.02) } dt = [(2B4 − 50B3 ) − 50B4 (t − 0.02)]

= 2 × 10−3

· e−25(t−0.02) × 10−3 A

υC1 (0) = υC1 (0− ) = 0, iC1 (0) = iC1 (0− ) = 0. Application of the expressions for B1 and B2 available in Table 6-1 gives B1 = υC (0) − υC (∞) = 0 − 12 = −12 V, 1 iC (0) + α[υC1 (0) − υC1 (∞)] C 1 = 0 + 25[0 − 12] = −300 V/s.

Across the juncture between time segment 1 and time segment 2, neither the voltage can change (as mandated by the capacitor) nor can the current (as mandated by the inductor). Thus,

(6.40a)

B2 =

(6.40b)

(6.45a)

iC1 (t = 20 ms) = iC2 (t = 20 ms).

(6.45b)

12 − (12 + 300 × 0.02)e−25×0.02 = B3 ,

υC1 (t) = 12 − (12 + 300t)e−25t V,

15 × 0.02e−25×0.02 = (2B4 − 50B3 ) × 10−3 ,

(6.41)

for 0 ≤ t ≤ 20 ms.

whose joint solution leads to B3 = 1.08 V,

The associated current is dυC1 d = 2 × 10−3 [12 − (12 + 300t)e−25t ] dt dt = 15te−25t A,

υC1 (t = 20 ms) = υC2 (t = 20 ms),

Application of Eqs. (6.45a and b) to the expressions given by Eqs. (6.41) to (6.44) gives

Consequently, υC1 (t) is given by

iC1 (t) = C

for t ≥ 20 ms. (6.44)

B4 = 118.04 V/s.

Consequently, υC2 (t) = [1.08 + 118.04(t − 0.02)]e−25(t−0.02) V

(for 0 ≤ t ≤ 20 ms). (6.42)

for t ≥ 20 ms

(6.46a)

and

Time Segment 2: t ≥ 20 ms. After moving the switch back to position 1 at t = 20 ms, the circuit no longer has any active sources, and yet it is part of a closed circuit (Fig. 6-12(c)), allowing the capacitor and inductor to dissipate their stored energies through the resistor. Hence, at t = ∞,

iC2 (t) = [0.182 − 5.90(t − 0.02)]e−25(t−0.02) A for t ≥ 20 ms.

(6.46b)

The waveforms of υC (t) and iC (t) are displayed in Figs. 6-12(d) and (e), respectively.

υC2 (∞) = 0. Upon shifting t by 0.02 s, the expression for υC2 (t) assumes the form υC2 (t) = [B3 + B4 (t − 0.02)]e

−25(t−0.02)

V

for t ≥ 20 ms,

(6.43)

Example 6-8: Two-Source Circuit

The switch in the circuit of Fig. 6-13(a) was opened at t = 0, after it had been closed for a long time. If Vs1 = 20 V, Vs2 = 24 V, R1 = 40 , R2 = R3 = 20 , R4 = 10 , L = 0.8 H, and C = 2 mF, determine υC (t) for t ≥ 0.

“book” — 2015/5/4 — 7:16 — page 352 — #23

352

CHAPTER 6

t=0

R1

R3

Vs2 +_

R1

R2 Vs1

+ _

C

iL

Vs2 +_

R3

R4

R2 Vs1

υC

+ _

RLC CIRCUITS

R4

I2

C iL(0−) = 0.2 A

I1

L

υC(0−) = −4 V

L

(a)

At t = 0−

(b) υC (V) 0

Vs2 +_

R3 R2

Req

a

iL

C

L

C

iL

0.4

0.6

0.8

1

t (s)

Underdamped response

−2 −3

b υC

0.2

−1

a iC

R4

b

Veq +_

0

υC

−4 −5

L

−6 −7 −8

At t > 0

(c)

(d) Equivalent circuit at t > 0

(e) υC(t)

Figure 6-13: Circuit for Example 6-8.

Solution: Consider the state of the circuit at t = 0− (before opening the switch), as depicted by Fig. 6-13(b). The mesh current equations for the indicated loops are −Vs1 + R1 I1 + R2 (I1 − I2 ) = 0,

Req = (R2 + R3 ) R4 =

R2 (I2 − I1 ) + R3 I2 + Vs2 + R4 I2 = 0. After substituting the given values for the sources and the resistors, simultaneous solution of the two equations leads to I1 = 0.2 A,

Next, we consider Fig. 6-13(c), which depicts the circuit configuration at t > 0 (after opening the switch). To simplify the analysis, we use source transformation to convert the circuit into its Th´evenin equivalent, as shown in Fig. 6-13(d), where

Veq =

Hence,

Req 8 = = 5 Np/s, 2L 2 × 0.8 1 1 =√ ω0 = √ = 25 rad/s. LC 0.8 × 2 × 10−3 α=



iL (0 ) = I1 = 0.2 A.

Vs2 × Req = 4.8 V. R2 + R 3

Now we are ready to analyze the series RLC circuit of Fig. 6-13(d). To that end, we compute α and ω0 :

I2 = −0.4 A.

υC (0− ) = I2 R4 = −0.4 × 10 = −4 V,

(R2 + R3 )R4 = 8 , R2 + R 3 + R 4

(6.47a) (6.47b)

“book” — 2015/5/4 — 7:16 — page 353 — #24

6-7 THE PARALLEL RLC CIRCUIT

353

Since α < ω0 , the capacitor voltage υC will exhibit an underdamped oscillatory response of the form given by Eq. (6.35) as υC (t) = {e−αt [D1 cos ωd t + D2 sin ωd t]} + υC (∞), (6.48)

Vs u(t)

+ + __

υC(t)

(a)

where ωd =

  ω02 − α 2 = 252 − 52 = 24.5 rad/s.

Is u(t)

It is evident from the circuit in Fig. 6-13(d) that υC (∞) = −Veq = −4.8 V.

D1 = υC (0) − υC (∞) = −4 + 4.8 = 0.8 V, D2 =

υC(t)

Parallel RLC

RLC circuit shown in (a) is identical in form to that of the current iL (t) in the parallel RLC circuit in (b).

υC dυC + iL + C = Is . R dt

(6.49b)

(6.52)

Using υC = υL = L diL /dt, and rearranging terms, leads to

(6.50)

d 2 iL 1 diL Is 1 + + iL = , 2 dt RC dt LC LC

iL + a2 iL + b2 iL = c2 ,

6-7 The Parallel RLC Circuit

(6.53)

which can be rewritten in the abbreviated form

The waveform of υC (t) is displayed in Fig. 6-13(e).

(6.54)

where

Having completed our examination of the series RLC circuit [Fig. 6-14(a)], we now turn our attention to the parallel RLC circuit shown in Fig. 6-14(b). As we will see shortly, the current iL (t) flowing through the inductor in the parallel RLC circuit is characterized by a second-order differential equation identical in form to that for the voltage υC (t) across the capacitor of the series RLC circuit. Accordingly, we will take advantage of this correspondence between the series and parallel RLC circuits by adapting the solutions we obtained in the preceding section for the series circuit to the solutions we seek in this section for the parallel circuit. Application of KCL to the circuit in Fig. 6-14(b) gives for t ≥ 0.

C

L

Figure 6-14: The differential equation for υC (t) of the series

υC (t) = {−4.8 + e−5t [0.8 cos 24.5t − 3.92 sin 24.5t]} V,

iR + iL + iC = Is

R

iC

i(t)

When expressed in terms of υC (t), the voltage common to all three passive elements, Eq. (6.51) becomes

With all unknown quantities accounted for,

for t ≥ 0.

iR

+ _

(6.49a)

1 C iC (0) + α[υC (0) − υC (∞)]

ωd −100 + 5[−4 + 4.8] = −3.92 V. = 24.5

Series RLC

(b)

To determine D1 and D2 , we apply Eq. (6.37) with υC (0) = −4 V, iC (0) = −iL (0) = −0.2 A, and υC (∞) = −4.8 V,

L

R

a2 =

1 , RC

1 , LC

c2 =

Is . LC

(6.55)

Comparison of Eq. (6.54) with Eq. (6.5) for the capacitor voltage of the series RLC circuit reveals that the two differential equations are identical in form, albeit the constant coefficients have different expressions in the two cases. The overdamped, underdamped, and critically damped expressions for iL (t) are given in Table 6-1. Quantities s1 , s2 , ω0 , and ωd retain the same expressions given earlier, but α is now given by α=

(6.51)

b2 =

1 2RC

(parallel RLC).

(6.56)

“book” — 2015/5/4 — 7:16 — page 354 — #25

354

CHAPTER 6

Parallel RLC

V0 12 = = 0.2 A, R1 60 R1 R 2 R  = R1 R2 = = 20. R1 + R 2 I0 =

Overdamped (α > ω0 ) iL (t) = [A1 es1 t + A2 es2 t + iL (∞)],

(for t ≥ 0) (6.57a)

Critically damped (α = ω0 ) iL (t) = [(B1 + B2 t)e−αt + iL (∞)],

1 1 = = 50 Np/s, 2R  C 2 × 20 × 500 × 10−6 1 1 =√ ω0 = √ = 100 rad/s. LC 0.2 × 500 × 10−6

(for t ≥ 0) (6.57b)

iL (t) = [e−αt (D1 cos ωd t + D2 sin ωd t) + iL (∞)], (for t ≥ 0)

For the parallel RLC circuit in Fig. 6-15(d), the expressions for α and ω0 are given by α=

Underdamped (α < ω0 )

∗ More

RLC CIRCUITS

(6.57c)

Since α < ω0 , the circuit will exhibit an underdamped response with a damped natural frequency ωd given in Table 6-1 as   ωd = ω02 − α 2 = 1002 − 502 = 86.6 rad/s. From Table 6-1, the expression for iL (t) is given by

details in Table 6-1.

iL (t) = [e−αt (D1 cos ωd t + D2 sin ωd t) + iL (∞)] for t ≥ 0.

Example 6-9: Parallel RLC Circuit

Determine iL (t) in the circuit of Fig. 6-15(a) for t ≥ 0, given that Is = 0.5 A, V0 = 12 V, R1 = 60 , R2 = 30 , L = 0.2 H, and C = 500 μF. Solution: The circuit in Fig. 6-15(b) represents the steady state condition of the circuit at t = 0− (prior to moving the switch). Under constant conditions, C acts like an open circuit and L acts like a short circuit. Given that Is flows entirely through the short circuit representing the inductor, it follows that iL (0− ) = Is = 0.5 A, υC (0− ) = 0. Since iL through an inductor cannot change instantaneously, nor can υC across a capacitor, these conditions are equally applicable at t = 0. Consequently, iL (0) = iL (0− ) = 0.5 A, and υL (0) = υC (0) = 0. After moving the switch (t > 0), the circuit assumes the configuration shown in Fig. 6-15(c). After application of source transformation, current source I0 and the equivalent resistance R  in Fig. 6-15(d) are given by

At t = ∞, the inductor behaves like a short circuit, forcing I0 to flow through it exclusively. Hence, iL (∞) = I0 = 0.2 A. The only remaining unknowns are D1 and D2 , which we determine by applying the expressions given in Table 6-1, namely D1 = iL (0) − iL (∞) = (0.5 − 0.2) A = 0.3 A, and υL (0) + α[iL (0) − iL (∞)] ωd 0 + 50(0.5 − 0.2) = = 0.17 A. 86.6

D2 =

1 L

The final expression for iL (t) is then given by iL (t) = [0.2 + e−50t (0.3 cos 86.6t + 0.17 sin 86.6t)] A, for t ≥ 0, and its plot is displayed in Fig. 6-15(e). Exercise 6-7: Determine the initial and final values for iL

in the circuit of Fig. E6.7 on the following page and provide an expression for iL (t).

“book” — 2015/5/4 — 7:16 — page 355 — #26

6-7 THE PARALLEL RLC CIRCUIT

355

υC(0−) iC

1

R1

Is

V0

iR

t=0

2

iL

R2

L

C

iL(0−)

υC

Is

R2

(b)

At t = 0−

L

(a)

iL

iL R1 V0

R2

L

C

I0

υC

R

L

C

υC

+ _ After t = 0

(c)

C

+ _

(d)

Norton equivalent

iL (A) 0.5

0.4

0.3

0.2 t (s) 0

0.05

0.1

0.15

(e) Plot of iL(t)

Figure 6-15: Circuit for Example 6-9.

2H

+ υL _ 15 mA

40 Ω

5 mF

iL

Answer: iL (0) = 5 mA, υL (0) = 0.4 V,

t=0

80 Ω

iL (∞) = 15 mA, α = 2.5 Np/s, ω0 = 10 rad/s, ωd = 9.68 rad/s, iL(t) = {15 − [10 cos 9.68t −18.08 sin 9.68t]e−2.5t } mA.

(See Figure E6.7

)

“book” — 2015/5/4 — 7:16 — page 356 — #27

356

TECHNOLOGY BRIEF 16: RFID TAGS AND ANTENNA DESIGN

Technology Brief 16 RFID Tags and Antenna Design RFID Applications Radio-frequency identification (RFID) uses electromagnetic fields to transfer identifying information from a small electrical ID circuit to an external receiver. These are commonly used for identifying or tracking animals, packages and goods, smart cards, tags, etc. (Fig. TF16-1). RFID circuits are injected in pets to help identify and return lost or stolen animals, attached via ear tags to livestock to identify their whereabouts and activities (how much time they spend eating or drinking), attached to athletes via wrist bands to track and verify their progress in a race, affixed to consumer goods and packaging to track, locate, and maintain inventory, and prevent theft. RFID tags can be based on either static, unchanging data (such as the ID number for a dog or cat), or their data can be changed by either an internal circuit (monitoring and reporting temperature of a refrigerated shipping container, for instance) or an external circuit (such as marking the last time a box was inspected).

When combined with other circuits, the information provided by RFID tags can be used in a myriad of ways. For instance, credit-card sized RFID tags attached to valuable art or other one-of-a-kind objects contain a unique ID number, as well as circuits detecting tilt and vibration. This information is continuously transmitted to receivers on the ceiling of a museum to create a security system that constantly monitors their location and status, and generates alarms if they are moved. RFID tags permanently installed in new guitars can help track them throughout their lives, and those installed in vintage guitars can help prevent fraud and theft. RFID tags are in most access-monitoring cards today, and can uniquely identify a person and his/her time of entry and exit. If other items are also tracked (sensitive documents for instance), an RFID reader can also identify what he/she is carrying and can generate an alarm if documents are leaving a room (or books leaving a library) that shouldn’t be. RFID tags can be used in numerous medical applications to identify a person and identify and track the drugs or treatments he/she receives. RFID and bar code scanners can be used for similar applications, but work in very different ways. Bar code scanners require direct visual access for a laser to read

11.5 mm

11.5 mm

Grain of rice

Figure TF16-1: RFID examples.

“book” — 2015/5/4 — 7:16 — page 357 — #28

TECHNOLOGY BRIEF 16: RFID TAGS AND ANTENNA DESIGN

357

RFID reader

Chip Antenna

Antenna

Transponder

Tag Figure TF16-2: RFID system.

the bar code. RFID circuits can be out of sight (inside a pet or package) as long as the wireless electromagnetic signal can penetrate the external packaging. Bar codes are read only. RFID systems can be read only or readwrite. Bar codes are printed directly on packaging, or stickers affixed to packaging. RFID systems require an external antenna and a (tiny) computer chip. The antenna can be printed, but the chip must be somehow affixed.The entire system is often implemented in a sticker or card. Bar codes are essentially free (printed), whereas RFID tags cost 15 US cents and up.

RFID Operation In a passive RFID system, an external transponder transmits a wireless signal to the RFID circuit (Fig. TF16-2), which “wakes up” and receives power from the signal through inductive coupling or other power harvesting methods. It then transmits its coded ID information back to the transponder, through the inductive link.The advantage of passive RFID systems is that they can be very small, not much bigger than a grain of rice, and can last for decades without maintenance as they do not require an internal battery to power the circuit. But the transponder must be within a short distance (less than 1 m) of the RFID circuit in order to receive the ID information. Active RFID systems have a battery to power the internal RFID circuit and can therefore transmit much further, up to 200 m. RFID systems consist of an RFID transceiver with a sinusoidal source and (typically) a loop antenna, through which the current flows, creating a magnetic field. The magnetic field is part of an electromagnetic wave that travels a short distance through the air to the RFID tag. The RFID tag has another (typically) loop or loop-like antenna to receive the magnetic field and convert it back to a current, and an RF circuit to convert it to a small

voltage that can be used to power the data circuit in the chip. Frequencies used for RFID and some of their applications are listed in Table TT16-1.

RFID Antennas Two examples of RFID antennas are shown in Fig. TF16-3. Both are printed 2-D antennas containing an inductor, in either a coiled design as in part (a) or in a “squiggly” design (yes, it really is called a squiggle tag),

Chip Substrate

Antenna coil

(a) Texas Instruments RFID tag

Chip

(b) Squiggle antenna Figure TF16-3: RFID antennas.

“book” — 2015/5/4 — 7:16 — page 358 — #29

358

TECHNOLOGY BRIEF 16: RFID TAGS AND ANTENNA DESIGN

Table TT16-1: RFID frequency bands. Approximate Tag Cost in Volume (2006) US$

Band

Regulations

Range

Data Speed

Remarks

120–150 kHz (LF)

Unregulated

10 cm

Low

Animal identification, factory data collection

$1

13.56 MHz (HF)

(ISM) band worldwide

10 cm – 1 m

Low to moderate

Smart cards (MIFARE, ISO/IEC 14443)

$0.50

433 MHz (UHF)

Short-range devices

1–100 m

Moderate

Defense applications, with active tags

$5

865–868 MHz (Europe), 902–928 MHz (North America) UHF

ISM band

1–12 m

Moderate to high

EAN, various standards

$0.15 (passive tags)

2450–5800 MHz (microwave)

ISM band

1–2 m

High

802.11 (WLAN), Bluetooth standards

$25 (active tag)

3.1–10 GHz (microwave)

Ultra wide band

1 to 200 m

High

Requires semi-active or active tags

$5

which is often printed on a sticker label for consumer products. Antenna design is a subspecialty of electrical engineering. Antenna designers consider ways to either convert current and voltage to electric and magnetic fields in the air (for wireless transmission) or to collect those fields in the air and convert them back into currents and voltages. In general, the same antenna can be used to receive and transmit the RFID signals. Antenna performance is governed by the shape of the antenna and its size relative to the wavelength λ of the electromagnetic (EM) wave it radiates or intercepts. The wavelength, in turn, is related to the signal frequency f by λ = c/f , where c is the velocity of light in vacuum. Hence, the size of an antenna usually is chosen to match the EM frequency that the RFID is intended to use. The ratio of electric to magnetic field is called the impedance of the antenna, and it needs to be matched to the same ratio of voltage and current that are produced or received by the circuit (the impedance of the circuit). The impedance of the circuit is controlled by the capacitors, resistors,

inductors, and other elements at the input or output of the circuit. The impedance of the antenna is controlled by its shape and size. Coils tend to be more inductive, which means their impedance is more like an inductor (has a positive imaginary part). Antennas shaped like plates tend to be more capacitive (having a negative imaginary part). Most antennas are a combination of inductive and capacitive, and can be modeled in circuit analysis as circuits containing both inductors and capacitors. Circuit elements are called lumped elements because their capacitance, inductance, and resistance are built from individual components, whereas an antenna is a distributed element whose capacitance, inductance, and resistance are spatially distributed along the length of the antenna.Taking all of these design factors into account at once is fairly daunting, so computer software is used extensively in antenna design, leading to creative designs such as the squiggle antenna and beyond. Antenna designers sometimes say they are “painting with copper” to describe the creative artistry of their field.

“book” — 2015/5/4 — 7:16 — page 359 — #30

6-8

GENERAL SOLUTION FOR ANY SECOND-ORDER CIRCUIT WITH DC SOURCES

Exercise 6-8: In the parallel RLC circuit shown in

Fig. 6-14(b), how much energy will be stored in L and C at t = ∞? Answer: wL = 21 LIs2 , wC = 0. (See

)

Step 1: Develop a second-order differential equation for x(t), for t ≥ 0. Express the equation in the general form x  + ax  + bx = c,

Step 2: Determine the values of α and ω0 : √ a α= , ω0 = b . 2

According to the material covered in the preceding sections, series and parallel RLC circuit share a common set of characteristics. An RLC circuit is characterized by a resonant frequency ω0 and a damping coefficient α, and when driven by a sudden dc excitation, the circuit exhibits a response that decays exponentially as e−αt , and it may or may not contain an oscillatory variation, depending on whether ω0 is or is not larger than α in magnitude, respectively. These characteristics arise from the interplay between energy storage and energy dissipation. During the operation of the RLC circuit, energy is exchanged between the two storage elements—the capacitor and the inductor—through the resistor. Dissipation is governed by e−αt , which we can redefine as e−t/τ , with

General Solution Overdamped (α > ω0 ) x(t) = [A1 es1 t + A2 es2 t + x(∞)],

(for t ≥ 0) (6.61a)

Critically Damped (α = ω0 )

(6.58)

In this alternative form, the decay rate is specified by the time constant τ . If τ is short (rapid decay) in comparison with the duration of a single oscillation period T , where T = 2π/ωd , it means that energy burns away too quickly to generate an oscillation. This is the overdamped case. On the other hand, if τ is sufficiently long (slow decay) in comparison with T , energy will move back and forth between L and C, generating an oscillation. With every cycle, however, the resistance will burn off some of the remaining energy, resulting in an underdamped response that decays and oscillates simultaneously. If R = 0, the circuit will oscillate forever at the resonant frequency ω0 (see Exercise 6-9). Building on the experience we gained from our examination of the series and parallel RLC circuits, we now extend the method of solution to any second-order circuit, including those containing op amps. For a circuit containing only dc sources (or no independent sources at all), we seek to find the circuit response x(t) for t ≥ 0, where x(t) is a voltage or current of interest in the circuit, and t = 0 is the instant at which the circuit experiences a sudden change (usually caused by a switch). To that end, we propose the following solution outline:

(6.60)

Step 3: Determine whether the response x(t) is overdamped, critically damped, or underdamped, and write down the expression corresponding to that case from the following general solution:

x(t) = [(B1 + B2 t)e−αt + x(∞)], (s).

(6.59)

where a, b, and c are constants.

6-8 General Solution for Any Second-Order Circuit with dc Sources

1 τ= α

359

(for t ≥ 0) (6.61b)

Underdamped (α < ω0 ) x(t) = [e−αt (D1 cos ωd t + D2 sin ωd t) + x(∞)], (for t ≥ 0) where s1 = −α +



(6.61c)

α 2 − ω02 ,

(6.62a)

s2 = −α − α 2 − ω02 ,  ωd = ω02 − α 2 .

(6.62b)



(6.62c)

 The three expressions given by Eq. (6.61) represent the circuit response to a sudden change that occurs at t = 0. Had the sudden change occurred at t = T0 instead, the expressions would continue to apply, but t will need to be replaced with (t − T0 ) everywhere on the right-hand side (only) of those expressions. 

“book” — 2015/5/4 — 7:16 — page 360 — #31

360

CHAPTER 6

RLC CIRCUITS

Table 6-2: General solution for second-order circuits for t ≥ 0. x(t) = unknown variable (voltage or current) Differential equation: x  + ax  + bx = c Initial conditions: x(0) and x  (0) c Final condition: x(∞) = b √ a α= ω0 = b 2

Overdamped Response α > ω0

x(t) = [A1 es1 t + A2 es2 t + x(∞)] u(t)  s1 = −α + α 2 − ω02 s2 = −α − α 2 − ω02    x  (0) − s2 [x(0) − x(∞)] x (0) − s1 [x(0) − x(∞)] A1 = A2 = − s1 − s 2 s1 − s 2 

Critically Damped α = ω0

x(t) = [(B1 + B2 t)e−αt + x(∞)] u(t) B1 = x(0) − x(∞) B2 = x  (0) + α[x(0) − x(∞)]

Underdamped α < ω0 D1 = x(0) − x(∞)

x  (0) + α[x(0) − x(∞)] D2 = ωd  ωd = ω02 − α 2

Step 4: Evaluate the circuit to determine x(∞) at t = ∞. Alternatively, we can use x(∞) =

c . b

(6.63)

Step 5: Apply initial conditions for x(t) and x  (t) at t = 0 (or at t = T0 if the sudden change occurred at T0 ) to determine the remaining unknown constants.

Step 1: Obtain differential equation for iL (t) After closing the switch, node 1 gets connected to node 2 and R2 becomes inconsequential to the rest of the circuit because it is connected in parallel with a short circuit. At node 2 of the circuit in Fig. 6-16(c), KCL gives −i1 + iL + iC = 0. In terms of the node voltage υC ,

This procedure is highlighted in Table 6-2 and demonstrated through Examples 6-10 to 6-12.

υC − V0 , R1 dυC iC = C . dt

−i1 =

Example 6-10: RLC Circuit with a Short-Circuit Switch

The switch in the circuit of Fig. 6-16(a) had been open for a long time before it was closed at t = 0. Determine iL (t) for t ≥ 0. The circuit elements have the following values: V0 = 24 V, R1 = 4 , R2 = 8 , R3 = 12 , L = 2 H, and C = 0.2 F. Solution: Figures 6-16(b), (c), and (d) depict the state of the circuit at t = 0− , t ≥ 0, and t = ∞, respectively.

(6.64)

(6.65a) (6.65b)

Hence, V0 dυC υC = + iL + C . (6.66) R1 dt R1 The voltage υC is equal to the sum of the voltages across L and R3 , diL (6.67) + i L R3 . υC = L dt

“book” — 2015/5/4 — 7:16 — page 361 — #32

GENERAL SOLUTION FOR ANY SECOND-ORDER CIRCUIT WITH DC SOURCES

R1

L + + __

V0

iC

iL

i1

C

i1

υC

8

i1( ) iC

L

iC( ) = 0

iL( ) 8

R1

R1 + + __

υC(0−) = 12 V

R3

υ2 = υC 2

i1

C

(b) At t = 0−: iL(0−) = V0 /(R1 + R2 + R3) = 1 A, and υC(0−) = iL(0−) R3 = 12 V.

iL

V0

L + + __

V0

R3

iC(0−) = 0

iL(0−) = 1 A

R1

(a) Circuit with switch

1

R2

1

2

8

R2

υ2(0−) 2

C

L

υC V0

R3

(c) At t > 0

+ + __

C

υC( ) 8

1

361

R3

(d) At t =

: iL( ) = V0 /(R1 + R3) = 1.5 A. 8

t=0

8

6-8

Figure 6-16: Circuit for Example 6-10.

where

Substituting Eq. (6.67) in Eq. (6.66) leads to 

L + R 1 R3 C 2 + 4 × 12 × 0.2 = 7.25, = R1 LC 4 × 2 × 0.2 R 1 + R3 4 + 12 b= = = 10, R1 LC 4 × 2 × 0.2 24 V0 = = 15. c= R1 LC 4 × 2 × 0.2

a=

  1 diL V0 d diL L . + i L R3 + i L + C L + i L R3 = R1 dt dt dt R1 (6.68) After carrying out the differentiation in the third term and rearranging terms, we have

d 2 iL + dt 2





L + R 1 R3 C R1 LC



diL + dt



R1 + R 3 R1 LC

iL =

+ biL = c,

(6.71c)

Step 2: Determine α and ω0 α= and

+ aiL

(6.71b)



V0 . R1 LC (6.69) For convenience, we rewrite Eq. (6.69) in the compact form

iL

(6.71a)

(6.70)

ω0 =

a 7.25 = 3.625 = 2 2 √ √ b = 10 = 3.162.

(6.72a)

(6.72b)

“book” — 2015/5/4 — 7:16 — page 362 — #33

362

CHAPTER 6 Since υL = L diL /dt, it follows that

Step 3: Determine damping condition and select appropriate expression Since α > ω0 , the response is overdamped, and iL (t) = A1 es1 t + A2 es2 t + iL (∞)

iL (0) = 0.

and s2 = −α −

 

(6.79)

(6.73)

with s1 = −α +

RLC CIRCUITS

α 2 − ω02 = −1.85 Np/s

(6.74a)

α 2 − ω02 = −5.40 Np/s.

(6.74b)

The expressions for A1 and A2 in Table 6-2 are given in terms of x, the variable associated with the second-order differential equation. In the present case, our differential equation is given by Eq. (6.70), with iL (t) as the unknown variable. Hence, by setting x = iL in the expressions for A1 and A2 , we have iL (0) − s2 [iL (0) − iL (∞)] s1 − s 2 0 + 5.4(1 − 1.5) = = −0.76 A −1.85 + 5.4

A1 = Step 4: Determine iL (∞) From the circuit in Fig. 6-16(d), iC = 0 (open-circuit capacitor) and iL (∞) =

V0 24 = 1.5 A. = R1 + R 3 4 + 12

(6.75)

and

(6.80a)

 iL (0) − s1 [iL (0) − iL (∞)] A2 = − s1 − s 2   0 + 1.85(1 − 1.5) =− = 0.26 A, −1.85 + 5.4 

(6.80b) (6.80c)

Step 5: Invoke initial conditions With C acting like an open circuit at t = 0− (Fig. 6-16(b)), IL (0− ) = i1 (0− ) =

V0 = 1 A. R1 + R 2 + R 3

and the final solution is then given by iL (t) = [1.5−0.76e−1.85t +0.26e−5.4t ] A

for t ≥ 0. (6.81)

Since iL cannot change in zero time, iL (0) = iL (0− ) = 1 A.

(6.76)

Exercise 6-9: Develop an expression for iC (t) in the

circuit of Fig. E6.9 for t ≥ 0.

We need one additional relationship involving A1 and A2 , which can be provided by the initial condition on iL . From the circuit in Fig. 6-16(b) at t = 0− , we have υC (0− ) = iL (0− ) R3 = 1 × 12 = 12 V.

(6.77)

As we transition from t = 0− (before closing the switch) to t = 0 (after closing the switch), neither iL nor υC can change, which means that the voltage υ2 (0) at node 2 will continue to be 12 V and the current iL through R3 will continue to be 1 A. Hence, the voltage υL (0) has to be υL (0) = υ2 (0) − iL (0) R3 = 12 − 1 × 12 = 0.

(6.78)

I0

+ _

iL

t=0 L

iC C

Figure E6.9



Answer: iC (t) = I0 cos ω0 t, with ω0 = 1/ LC . This

is an LC oscillator circuit in which dc energy provided by the current source is converted into ac energy in the LC circuit. (See )

“book” — 2015/5/4 — 7:16 — page 363 — #34

TECHNOLOGY BRIEF 17: NEURAL STIMULATION AND RECORDING

Technology Brief 17 Neural Stimulation and Recording Section 4-12 introduced neural probes and how they can be used to measure voltage at specific locations in the brain. They can also be used to stimulate neurons to control movement, sight, hearing, touch, smell, emotion, and more. Neural stimulation and recording begin with a neural probe such as the three dimensional neural probe shown in Fig. 4-30 or the spiral-shaped cochlear implant electrodes shown in Fig. TF17-1. Each electrode is meant to stimulate one or more nearby neurons. The electrodes are surgically inserted in proximity to the neurons of interest, and connected onto an electrical stimulation device that sends carefully designed electrical pulses into the extracellular fluid around them (for neural stimulation), or connected to an electrical receiver (that reads signals from them in the case of neural recording). There are many different devices, both commercially available and in research applications, that utilize neural stimulation or recording. These bioelectronics are one of the most exciting and rapidly advancing areas of electrical engineering. Several examples of these devices are given below.

363

by a microphone and electrical circuitry. The sounds are picked up by the microphone mounted behind the ear, processed or coded (using electrical circuitry) into electrical pulses associated with the sounds, and then transmitted through the skin via inductive coupling or direct connection to the electrodes. The electrodes place these signals directly onto the auditory nerves, which then send the signals to the brain, which “hears” the sound. If the auditory nerve is not functional, an auditory brainstem implant is used instead, wherein electrodes directly stimulate the cochlear nucleus complex in the lower brain stem.

Artificial Eye Retina The artificial retina, or cortical implant, replaces damaged eye structures with an external camera, a wireless link (shown as the two orange inductive coils in Fig. TF17-3), and an electrode array that stimulates the optic nerve in the back of the eye. Another alternative is to bypass the optical nerve and stimulate the visual cortex of the brain directly. The resolution of sight depends on the number of electrodes, as shown in Fig. TF17-4.

Brain Stimulation Cochlear Implant In the cochlear implant shown in Fig. TF17-2, the ear drum and stapes (inner bones of the ear) are replaced

Electrodes

Figure TF17-1: Preformed spiral electrode for cochlear c 2015 implant. (Courtesy of Cochlear Americas, Cochlear Americas.)

The deep brain stimulation (DBS) or cognitive prosthesis shown in Fig. TF17-5 is used to stimulate

1. Sounds are picked up by the microphone. 2. The signal is then “coded” (turned into a special pattern of electrical pulses). 3. These pulses are sent to the coil and are then transmitted across the skin to the implant. 4. The implant sends a pattern of electrical pulses to the electrodes in the cochlea. 5. The auditory nerve picks up these electrical pulses and sends them to the brain. The brain recognizes these signals as sound.

Figure TF17-2: A cochlear implant stimulates the auditory nerves to help deaf people hear. (Courtesy MEDEL.)

“book” — 2015/5/4 — 7:16 — page 364 — #35

364

TECHNOLOGY BRIEF 17: NEURAL STIMULATION AND RECORDING

Figure TF17-3: Artificial retina simulates the optic nerve to help blind people see. (Credit: John Wyatt.)

nerves deep within the brain. This has been used to reduce tremors due to Parkinson’s disease and to relieve some types of depression, and it has been proposed for treating a number of other psychological and physiological disorders. The development of applications for direct stimulation of the brain is often preceded by neural recording, to help researchers better understand the natural electrical signals in the body.

Figure TF17-5: Deep brain stimulation (DBS) is used to treat depression and tremors associated with Parkinson’s disease. (Credit: Medtronic.)

them), thus returning some level of motion control. If a limb is entirely gone, it can be replaced by an artificial limb, controlled by neural recording and stimulation (Fig. TF17-6). An interesting phenomenon associated with these and many other types of neural prosthetics is that the plasticity of the brain often allows the user to learn and train the brain and body to see, hear, touch, and move based on the adapted machine-brain interface from the neural signals.

Sensory and Motor Prostheses Several designs of sensory/motor prostheses are being developed to help patients with spinal cord injuries, damaged or amputated limbs, loss of bladder control, and other physical impairments. If only the nerve connections are damaged, these may be replaced by neural recording (to receive signals) and stimulation devices (to transmit

16 electrodes

200+ electrodes

1000+ electrodes

Figure TF17-4: Vision resolution expected with various numbers of sight-stimulating electrodes.

Pain Control Another application of both internal and external electrical stimulation is in control of pain. Basically, the pain signals are masked by a stimulation-induced tingling known as paresthesia. Internal devices used to induce paresthesia include the spinal cord stimulator (SCS) shown in Fig.TF17-7 and external devices include pulsed electromagnetic field (PEMF) stimulators. External devices use one of two methods for directing the pulsed energy to the location of the pain. One method involves inductive coupling (using coils external to the body), and the other involves the use of two electrodes on either side of the region, transmitting current from one electrode through the body region to the other electrode (Fig. TF17-8). PEMF devices have also been used to improve bone and soft tissue healing. Emerging technology in neural prostheses and other body-machine interfaces has already provided life improvements for many. This technology is still in its infancy,

“book” — 2015/5/4 — 7:16 — page 365 — #36

TECHNOLOGY BRIEF 17: NEURAL STIMULATION AND RECORDING

365

Figure TF17-6: Mind-controlled bionic arm uses both neural recording and neural stimulation within the brain and at the attachment site of the artificial limb. (Credit: Todd Kuiken, MD, Center for Bionic Medicine.)

Figure TF17-7: Spinal cord stimulator (SCS). (Credit: Spine-health.com.)

and many interesting challenges remain. How to create a full-function, long-term biocompatible implant small enough to be placed directly into the eye, brain, spine, bladder, brain and other organs, with battery life and/or power harvesting to support its operation, but with heat and power low enough not to damage the critical neurons it is connected to, surgically placing it correctly every time

Figure TF17-8: Wearable pulsed electromagnetic field (PEMF) pain-control device for the knee. (Credit: Orthomedical.)

for every patient, with easy ways to get information to and from the device . . . there are enough challenges to keep engineers engaged for decades to come!

“book” — 2015/5/4 — 7:16 — page 366 — #37

366

CHAPTER 6

R1 Vs u(t)

R2 i2

i1

+ _

ix

iy

L1

L2

(a) Circuit R1

R2 i2( )

8

8

i1( ) iy( )

L1

8

ix( ) 8

Vs

+ _

L2

RLC CIRCUITS

To obtain an expression for iy , we simply take the derivative of Eq. (6.85), R1  iy = i + ix . (6.86) L1 x After inserting Eqs. (6.85) and (6.86) into Eq. (6.84) and rearranging terms, we have     (R1 + R2 )L1 + R1 L2  R 1 R2 R2 Vs ix + ix = , ix + L1 L2 L1 L 2 L 1 L2 (6.87) which can be rewritten in the compact form ix + aix + bix = c,

(6.88)

where

(R1 + R2 )L1 + R1 L2 = 7.5, L1 L2 R1 R2 R 2 Vs b= = 6, c= = 21. L1 L2 L 1 L2 Step 2: Evaluate α, ω0 , s1 , and s2

(b) At t =

8

a=

Figure 6-17: Circuit for Example 6-11.

a 7.5 = = 3.75 Np/s, 2 2 √ √ ω0 = b = 6 = 2.45 rad/s,  s1 = −α + α 2 + ω02  = −3.75 + (3.75)2 − 6 = −0.91 Np/s, α=

Example 6-11: Two-Inductor Circuit

Determine i1 (t) and i2 (t) in the circuit of Fig. 6-17 for t ≥ 0. The component values are Vs = 1.4 V, R1 = 0.4 , R2 = 0.3 , L1 = 0.1 H, and L2 = 0.2 H. Solution: We designate ix and iy as the mesh currents in the two loops, as shown. We will analyze the circuit in terms of ix and iy and then use the solutions to determine i1 and i2 . For t ≥ 0, the mesh equations are given by: d (ix − iy ) = 0, dt diy d L1 (iy − ix ) + R2 iy + L2 = 0, dt dt −Vs + R1 ix + L1

−L1 ix + R2 iy + (L1 + L2 )iy = 0.

(6.89d)

Since α > ω0 , ix will exhibit an overdamped response given by ix (t) = [ix (∞) + A1 es1 t + A2 es2 t ]

(6.82)

(iy loop)

(6.83)

Take the time derivative of all terms in the iy -loop equation: (6.84)

To convert Eq. (6.84) into a differential equation in ix alone, we need to develop expressions for iy and iy in terms of ix and its derivatives. By isolating iy in Eq. (6.82), we have R1 Vs = ix + ix − . L1 L1

(3.75)2 − 6 = −6.6 Np/s.

(iy loop)

Step 1: Develop a differential equation in ix alone

iy



(6.89c)

Step 3: Write expression for ix (t)

(ix loop)

−L1 ix + R2 iy + (L1 + L2 )iy = 0.

s2 = −3.75 −

(6.89b)

(ix loop)

which can be rearranged and rewritten in the form R1 ix + L1 ix − L1 iy = Vs ,

and

(6.89a)

(6.85)

= [ix (∞) + A1 e−0.91t + A2 e−6.6t ].

(6.90)

Step 4: Evaluate final condition At t = ∞, the inductors in the circuit behave like short circuits (Fig. 6-17(b)), in which case the current generated by Vs will flow entirely through L1 . Hence, Vs 1.4 = = 3.5 A (6.91a) ix (∞) = R1 0.4 and iy (∞) = 0.

(6.91b)

The expression for ix (t) becomes ix (t) = 3.5 + A1 e−0.91t + A2 e−6.6t .

(6.92)

“book” — 2015/5/4 — 7:16 — page 367 — #38

6-8

GENERAL SOLUTION FOR ANY SECOND-ORDER CIRCUIT WITH A DC SOURCE

using the iy -loop equation (Eq. (6.83)) to generate expressions for ix and ix . The procedure leads to

Step 5: Invoke initial conditions Before t = 0, the circuit contained no sources. Hence, i1 (0) = i1 (0− ) = 0

367

iy (t) = 1.23(e−0.91t − e−6.6t ) A.

(6.93a)

(6.99)

Finally, the solutions for i1 (t) and i2 (t) are:

and i2 (0) = i2 (0− ) = 0,

i1 (t) = ix (t) − iy (t)

(6.93b)

= [3.5 − 1.59e−0.91t − 1.91e−6.6t ] A

which implies that ix (0) = ix (0− ) = 0

(6.94)

iy (0) = iy (0− ) = 0.

(6.95)

and

At t = 0, with no currents flowing through either loop, the voltages across L1 and L2 are both equal to Vs . That is, i1 (0) =

1 Vs υL (0) = L1 1 L1

and i2 (t) = iy (t) = 1.23(e−0.91t − e−6.6t ) A

Exercise 6-10: For the circuit in Fig. E6.10, determine

iC (t) for t ≥ 0.

(6.96a)

t=0 1 Vs υL2 (0) = , L2 L2

(6.100b)

(for t ≥ 0)

and i2 (0) =

(6.100a)

(6.96b)

iC 3Ω 3Ω

2A 2H

20 mF

Consequently, ix (0) = i1 (0) + i2 (0) =

Vs Vs + = 21. L1 L2

(6.97)

Now that we know the values of ix (0), ix (0), and ix (∞), we can apply the general expressions for A1 and A2 in Table 6-2 to get

)

Determine iL (t) in the op-amp circuit of Fig. 6-18(a) for t ≥ 0. Assume Vs = 1 mV, R1 = 10 k, R2 = 1 M, R3 = 100 , L = 5 H and C = 1 μF. Solution: KCL at node υn gives

 ix (0) − s1 [ix (0) − ix (∞)] A2 = − s1 − s 2   21 + 0.91(0 − 3.5) =− = −3.14 A. −0.91 + 6.6 

i1 + in + i2 + i3 = 0, or equivalently, υn − υout d υn − Vs + in + +C (υn − υout ) = 0. (6.101) R1 R2 dt

The final expression for ix (t) is then given by ix (t) = [3.5 − 0.36e−0.91t − 3.14e−6.6t ] A.

Answer: iC(t) = 2e−1.5t cos 4.77t A. (See

Example 6-12: Second-Order Op-Amp Circuit

i  (0) − s2 [ix (0) − ix (∞)] A1 = x s1 − s 2 21 + 6.6(0 − 3.5) = = −0.36 A −0.91 + 6.6 and

Figure E6.10

(6.98)

Repetition of steps 1–4 for iy requires that we start by taking the time derivative of the ix -loop equation (Eq. (6.82)) and then

Since υn = υp = 0, in = 0, and υout = R3 iL + L

diL , dt

(6.102)

“book” — 2015/5/4 — 7:16 — page 368 — #39

368

CHAPTER 6

R1 +

+ V u(t) -_ s

i2

i1

Rearranging, we have

C

i3

RLC CIRCUITS

iL + aiL + biL = c,

in

R2

where L + R2 R3 C = 21, R2 LC R3 b= = 20, R2 LC

_ υout

υn υp

+

(6.104)

a=

iL

R3 and

L

c=

Op-amp circuit

(a)

The damping behavior of iL is determined by how the magnitude of α compares with that of ω0 :

C

a = 10.5 Np/s, 2 √ √ ω0 = b = 20 = 4.47 rad/s.

R2

R1

α=

_ 8

υout( ) iL( ) 8

+ _ Vs

−Vs = −0.02. R1 LC

+

R3

Since α > ω0 , iL will exhibit an overdamped response given by iL (t) = [A1 es1 t + A2 es2 t + iL (∞)] u(t),

L 8

with

At t =

(b)

s1 = −α +

C

s2 = −α −

R2

R1

−_

+ + V -_ s

+

υout(0) iL(0)

 

α 2 − ω02 = −1.0, α 2 − ω02 = −20.

At t = ∞, the circuit assumes the equivalent configuration shown in Fig. 6-18(b), which is an inverting amplifier with an output voltage

R3

υout (∞) = −

L

R2 Vs . R1

Hence,

At t = 0

(c)

iL (∞) = Figure 6-18: Op-amp circuit of Example 6-12.

υout (∞) R 2 Vs =− = −1 mA. R3 R1 R3

The expression for iL (t) becomes iL (t) = [A1 e−t + A2 e−20t − 10−3 ].

Equation (6.101) becomes R3 iL + R2



L + R3 C R2



d 2 iL diL Vs + LC =− . (6.103) 2 dt dt R1

(6.105)

To determine the values of A1 and A2 , we examine initial conditions for iL and iL . At t = 0− , there were no active sources

“book” — 2015/5/4 — 7:16 — page 369 — #40

6-9

MULTISIM ANALYSIS OF CIRCUIT RESPONSE

in the circuit, and since iL cannot change instantaneously, it follows that iL (0) = iL (0− ) = 0, which means that the inductor behaves like an open circuit at t = 0, as depicted in Fig. 6-18(c). Also, since the voltage υC across the capacitor was zero before t = 0, it has to remain at zero at t = 0, which is why it has been replaced with a short circuit in Fig. 6-18(c). Consequently, υout (0) = 0, υL (0) = 0, and 1 iL (0) = υL (0) = 0. L From Table 6-2, with x = iL , iL (0) − s2 [iL (0) − iL (∞)] s1 − s 2 0 + 20(0 + 1) × 10−3 = 1.05 mA = −1 + 20

A1 =

and

(6.106)

 iL (0) − s1 [iL (0) − iL (∞)] s1 − s 2   0 + 1(0 + 1) =− × 10−3 = −0.053 mA. −1 + 20 

A2 = −

(6.107)

The final expression for iL (t) is then given by iL (t) = [1.05e−t − 0.053e−20t − 1] mA,

for t ≥ 0.

Concept Question 6-6: A circuit contains two capacitors

and three inductors, in addition to resistors and sources. Under what circumstance is it a secondorder circuit? (See )

369 with Multisim. As an example of a real-world application of the RLC-circuit response, we will then examine how such a circuit is used in RFID (radio frequency identification) technology.

6-9.1 The Series RLC Circuit Using the now (hopefully) familiar schematic tools, draw a series RLC circuit, including a switch, in the Multisim Schematic Capture window. Use the parts and component values listed in Table 6-3, and add an oscilloscope as shown in Fig. 6-19. The scope is used for both L1 and C1 , so that we may compare the voltages across them on the same screen. Make sure that before starting the interactive simulation, the initial condition of the switch is in position 2, so that the dc voltage source is not connected directly to the RLC circuit. Upon starting the simulation, you should see no voltage across any of the three components. After hitting the space bar to move the switch (Fig. 6-20), υL (t) will initially jump in level to 1 V and then exhibit an underdamped oscillatory response as a function of time. In contrast, υC (t) will exhibit an oscillatory behavior that will dampen out with time to assume a final value of 1 V. A note on the Interactive Simulation settings is appropriate here. When you run an Interactive Simulation, Multisim numerically solves for the solution to the circuit at successive points in time. The resolution of this time step can be modified under Simulate → Interactive Simulation Settings. Both the maximum time step (TMAX) and the initial time step can be changed. Normally, there is no reason to do this and Multisim’s defaults will work well. However, when using the virtual instruments, sometimes time points are generated too quickly and this makes it difficult for the user to observe the behavior, or conversely the resolution may be too small so that

Concept Question 6-7: Suppose a = 0 in Eq. (6.59). What type of response will x(t) have in that case? (See )

6-9 Multisim Analysis of Circuit Response Understanding the behavior of even a simple RLC circuit is sometimes a challenging task for electrical and computer engineering students. In reaction to a sudden change, a circuit gives rise to voltage and current variations that depend on the circuit topology, the initial conditions of its components, and the values of those components. In this section, we describe how to use Multisim to analyze the response of the series RLC circuit we discussed in earlier sections. The procedure is intended to demonstrate the steps one would follow to analyze any circuit

Position 1

Position 2

Figure 6-19: Multisim screen with RLC circuit.

“book” — 2015/5/4 — 7:16 — page 370 — #41

370

CHAPTER 6

RLC CIRCUITS

Table 6-3: Component values for the circuit in Fig. 6-19. Component

Group

Family

Quantity

Description

1

Basic

Resistor

1

1  resistor

300 m

Basic

Inductor

1

300 mH inductor

5.33 m

Basic

Capacitor

1

5.33 mF capacitor

SPDT

Basic

Switch

1

Single-pole double-throw (SPDT) switch

DC POWER

Sources

Power Sources

1

1 V dc source

Interactive Simulation

υC(t)

Exercise 6-12: Is the natural response for the circuit in

Fig. 6-19 over-, under-, or critically damped? You can determine this both graphically (from the oscilloscope) and mathematically, by comparing ω0 and α. Answer: (See

)

υL(t) Switch moved from position 2 to position 1

Exercise 6-13: Modify the value of R in the circuit of Fig. 6-19 so as to obtain a critically damped response. Answer: (See

)

Figure 6-20: Voltage responses to moving the switch in the RLC circuit from position 2 to position 1.

the progression of time in the Interactive Simulation becomes annoyingly slow. When generating the traces in Fig. 6-20, for example, it may be difficult to see the damped behavior directly on the scope window because it scrolls by too fast. In that case, it can be helpful to reduce both the maximum and initial time steps (10–100 × reduction usually works fine). This forces the computer to simulate more data points and slows it down, allowing you to see the trace appear more slowly. The drawback of this tweak is that you also use up more memory (and filespace). Exercise 6-11: Given the component values in the

Multisim circuit of Fig. 6-19, what are the values of ω0 and α for the circuit response? Answer: (See

)

6-9.2

RFID Circuit

Radio frequency identification (RFID) circuits are fast becoming ubiquitous in many mass consumer applications, ranging from tracking parcels and shipments to “smart” ID badges (see Technology Brief 16). Most systems in use today rely on a transceiver (usually handheld) that can remotely interrogate one or more RFID tags (ranging in size from a few millimeters to a few centimeters). Some tags reply with only a serial number, while others are connected to miniature sensors and return values for temperature, humidity, acceleration, position, etc. The key to the widespread success of these RFID tags is that they do not require batteries to operate! If the transceiver is in close proximity to the tag (usually within a fraction of a meter), the radio-frequency power it transmits is sufficient to activate the RFID tag. The RFID tag uses an RLC circuit to harvest this power and communicate back to the transceiver (Fig. 6-21). The essential elements of the RFID communication system are shown in the circuit of Fig. 6-22. [An actual RFID circuit is more sophisticated, but the basic principle

“book” — 2015/5/4 — 7:16 — page 371 — #42

6-9

MULTISIM ANALYSIS OF CIRCUIT RESPONSE

371

RF transceiver

er

ceiv

s Tran _

Vs

+ Antenna 1 (Ls)

Antenna 2 (Lp) RFID tag Rp Cp

Magnetic field coupling

Figure 6-21: Illustration of an RFID transceiver in close proximity to an RFID tag. Note that the RFID tag will only couple to the transceiver when the two inductors are aligned along the magnetic field (shown in blue).

To receiver circuits R T υout(t)

~+− υ

s

RFID transceiver

Ls

Magnetic field

Lp C p

υC

Rp

RFID tag

Figure 6-22: Basic elements of the RFID.

of operation is the same.] In transmit mode—with the SPDT switch connected to terminal T —the transceiver circuit consists

of a ac voltage source, υs , connected in series with inductor Ls . By moving the switch to terminal R, the transceiver circuit becomes a receiver with output voltage υout (t). In transmit mode, υs generates a current through Ls , which induces a magnetic field around it. If inductor Lp of the RFID tag is close to Ls , the magnetic field generated by Ls will induce a current through Lp . This current becomes the power source in the RFID-tag circuit, and the mechanism for building up the voltage across Cp to some maximum value VC . When the switch is moved from transmit mode to receive mode, υs stops delivering power to Ls . The current through Lp , however, cannot change to zero instantaneously. The RLC circuit will react to the sudden change with an oscillatory underdamped response characterized by a damped natural frequency ωd , whose value is governed by the choice of values for Rp , Lp , and Cp of the RFID tag. This oscillation frequency becomes part of the ID of that particular tag. In the same way

“book” — 2015/5/4 — 7:16 — page 372 — #43

372

CHAPTER 6

RLC CIRCUITS

Table 6-4: Parts for the Multisim circuit in Fig. 6-23. Component

Group

Family

TS IDEAL

Basic

Transformer

1

1 mH:1 mH ideal transformer

1k

Basic

Resistor

1

1 k resistor



Basic

Capacitor

1

1 μF capacitor

SPDT

Basic

Switch

1

SPDT switch

AC CURRENT

Sources

Signal Current Source

1

1 mA, 5.033 kHz

that magnetic coupling served to transfer power from Ls to Lp during the transmit mode, it also serves to transfer information in the opposite direction—from Lp to Ls —during the receive mode. Since diLs , υout (t) = Ls dt the output voltage recorded after moving the switch to receive

Quantity

Description

mode provides the reply by the RFID tag to the earlier excitation introduced by υs during the transmit mode. [Real RFID transceivers transmit a few bits of data by superimposing digital bits onto the oscillations.] To illustrate the operation of the RFID tag, we can simulate the process in Multisim. Using the parts listed in Table 6-4, we can build the circuit shown in the Multisim window of Fig. 6-23.

vout

R T

vout(t)

Figure 6-23: Multisim rendition of RFID circuit.

vC(t)

“book” — 2015/5/4 — 7:16 — page 373 — #44

6-9

MULTISIM ANALYSIS OF CIRCUIT RESPONSE

373

Concept Question 6-8: How does the transmitter in the

Channel A Voltage (V)

RFID system transfer power to the RLC circuit? (See ) Concept Question 6-9: How does the transceiver elicit a

reply from the RFID tag? (See

)

Exercise 6-14: Calculate ω0 , α, and ωd for the RLC circuit

Switch moved from T to R

in Fig. 6-23. How do ω0 and ωd compare with the angular frequency of the current source? This result, as we will learn later when we study resonant circuits in Chapter 9, is not at all by coincidence.

Answer: (See

Time (s) Figure 6-24: Oscilloscope trace for RFID receive channel υout (t) after moving the switch from T to R.

To simulate magnetic coupling between inductors Ls and Lp , we use transformer T1 , which represents two closely coupled inductors sharing a common magnetic field. In Multisim we set the inductance of each of the two transformer units to 1 mH and the coupling coefficient to 1. The circuit uses an oscilloscope to monitor υout (t). The oscilloscope trace is displayed in Fig. 6-24. Note that when the switch is moved from transmit to receive mode, υout (t) exhibits an immediate response that then decays exponentially with time. You may also want to plot υC (t) and iC (t) to examine the voltage and current experienced by the RFID tag itself during transmit and receive periods.

)

Exercise 6-15: Ideally, we would like the response of the

RFID tag to take a very long time to decay down to zero, so as to contain as many digital bits as possible. What determines the decay time? Change the values of some of the components in Fig. 6-23 so as to decrease the damping coefficient by a factor of 2. Answer: (See

)

Summary Concepts • Under dc steady state conditions, a capacitor behaves like an open circuit and an inductor behaves like a short circuit. • Second-order circuits include series and parallel RLC circuits, as well as any circuit containing two passive, energy storage elements (capacitors and inductors). • The response of a second-order circuit (containing dc sources) to a sudden change consists of a transient

component, which decays to zero as t → ∞, and a steady state component that has a constant value. • The transient response may be overdamped, critically damped, or underdamped, depending on the values of the circuit elements. • The general solution for second-order circuits is applicable to circuits containing op-amps. • Multisim can be used to simulate the response of any second-order circuit.

“book” — 2015/5/4 — 7:16 — page 374 — #45

374

CHAPTER 6

RLC CIRCUITS

Mathematical and Physical Models Step response of series and parallel RLC circuits (See Table 6-1)

General Solution for Second Order Circuits (cont’d.): Overdamped Response (α > ω0 ) x(t) = [x(∞) + A1 es1 t + A2 es2 t ] u(t)

General Solution for Second Order Circuits: (see details in Table 6-3) x  + ax  + bx = c

Differential equation

Important Terms

Critically Damped Response (α > ω0 ) x(t) = [x(∞) + (B1 + B2 t)e−αt ] u(t) Underdamped Response (α > ω0 ) x(t) = [x(∞) + [D1 cos ωd t + D2 sin ωd t]e−αt u(t)

Provide definitions or explain the meaning of the following terms:

characteristic equation critically damped critically damped response damped natural frequency damping coefficient initial condition initial time step

final condition first-order circuit homogeneous homogeneous solution invoke initial and final conditions MEMS maximum time step natural response

nepers/second oscillator overdamped response particular particular solution resonant frequency radio frequency identification RFID

second-order circuit steady-state steady-state response time constant time period transient transient response underdamped response

that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine (a) υC (0) and iL (0), (b) iC (0) and υL (0), and (c) υC (∞) and iL (∞).

PROBLEMS Section 6-1: Initial and Final Conditions *6.1 The SPST switch in the circuit of Fig. P6.1 closes at t = 0 after it had been open for a long time. Draw the configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine (a) υC (0) and iL (0), (b) iC (0) and υL (0), and (c) υC (∞) and iL (∞).

υL 2Ω

iL

L

iC

5Ω υL

iL

4Ω iC

L

t=0

C

υC



3Ω + 12 V _

+ 18 V _

t=0

C

υC Figure P6.2: Circuit for Problem 6.2.

Figure P6.1: Circuit for Problem 6.1.

6.2 The SPST switch in the circuit of Fig. P6.2 opens at t = 0, after it had been closed for a long time. Draw the configurations ∗ Answer(s) available in Appendix G.

6.3 The SPST switch in the circuit of Fig. P6.3 opens at t = 0, after it had been closed for a long time. Draw the configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine *(a) υC (0) and iL (0), (b) iC (0) and υL (0), and

“book” — 2015/5/4 — 7:16 — page 375 — #46

PROBLEMS

375

(c) υC (∞) and iL (∞).

iC

t=0 4A 2 kΩ

υL



iL

4 kΩ iC

+ 6V _

υC

C

iL

L

+ 10 V _

t=0

24 Ω

υL

L

υC

C + 12 V _

Figure P6.5: Circuit for Problems 6.5 and 6.6.

6.6 Repeat Problem 6.5, but start with a closed switch that opens at t = 0.

Figure P6.3: Circuit for Problem 6.3.

*6.7

6.4 The SPST switch in the circuit of Fig. P6.4 opens at t = 0, after it had been closed for a long time. Draw the configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine (a) υC (0) and iL (0), (b) iC (0) and υL (0), and (c) υC (∞) and iL (∞).

t=0



For the circuit in Fig. P6.7, determine i1 (0) and i2 (0).

1 2

8Ω i2

i1

t=0

+ 30 V _





L1

L2

5Ω Figure P6.7: Circuit for Problem 6.7.

10 Ω + 45 V _

8Ω iL

L

υL

iC C

υC

6.8 For the circuit of Fig. P6.8, determine (a) iC1 (0), iR1 (0), iC2 (0), and iR2 (0) and (b) υC1 (∞) and υC2 (∞).

3Ω Figure P6.4: Circuit for Problem 6.4.



6.5 The SPST switch in the circuit of Fig. P6.5 closes at t = 0, after it had been opened for a long time. Draw the configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine (a) υC (0) and iL (0), (b) iC (0) and υL (0), and (c) υC (∞) and iL (∞).

t=0 + _ 20 V

5Ω + 10 V _

iR1

C1 iC1 iR2

υC1 2Ω

C2

Figure P6.8: Circuit for Problem 6.8.

iC2 υC2

“book” — 2015/5/4 — 7:16 — page 376 — #47

376

CHAPTER 6

6.9 For the circuit in Fig. P6.9: (a) Draw the configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞. (b) Use the configurations to determine iL (0− ), υC (0− ), iL (0), υC (0), iL (∞), and υC (∞).



+ i1

30 V

t=0

t=0

+ _



t=0

+ _

+ _ υC

C



υC _

i2 3Ω

i3

+

iL 5V

RLC CIRCUITS

υL

6Ω 5A

L

_



10 Ω Figure P6.11: Circuit for Problem 6.11. Figure P6.9: Circuit for Problem 6.9.

*6.10 For the circuit in Fig. P6.10, determine iC (0), υC (0), iR (0), υR (0), iL (0), υL (0), υL (∞), iR (∞), υC (∞), and iL (∞).

12 Ω

_ υC

iC

iL

+ iR

+ _

24 V

t=0

+ L υ _R



+ _

υL

6.13 Determine iL (t) in the circuit of Fig. P6.12 and plot its waveform for t ≥ 0, given that V0 = 12 V, R1 = 0.4 , R2 = 1.2 , L = 0.1 H, and C = 0.1 F. Use a time scale that appropriately captures the shape of the waveform in your plot. *6.14 In the circuit of Fig. P6.12, V0 = 12 V, R1 = 0.4 , R2 = 1.2 , and L = 0.1 H. What should the value of C be in order for iL (t) to exhibit a critically damped response? Provide an expression for iL (t) and plot its waveform for t ≥ 0. 6.15 The voltage υ in a certain circuit is described by the differential equation

Figure P6.10: Circuit for Problem 6.10.

6.11 For the circuit in Fig. P6.11, find i1 (0− ), i2 (0), υC (0), and i3 (∞).

(a) Determine the values of α and ω0 .

Sections 6-2 to 6-6: Series RLC Circuit *6.12 Determine υC (t) in the circuit of Fig. P6.12 and plot its waveform for t ≥ 0, given that V0 = 12 V, R1 = 0.4 , R2 = 1.2 , L = 0.1 H, and C = 0.4 F. Use a time scale that appropriately captures the shape of the waveform in your plot.

L

R1 + V0 _

t=0

R2

3υ  + 24υ  + 75υ = 0.

(b) What type of damping is exhibited by υ(t)? *6.16 In the circuit of Fig. P6.16, the switch is moved from position 1 to position 2 at t = 0. Provide an expression for υC (t) for t ≥ 0.

iL C

Figure P6.12: Circuit for Problems 6.12 to 6.14.

R υC

+ V0 _

L

1 2

t=0

C

Figure P6.16: Circuit for Problem 6.16.

υC

“book” — 2015/5/4 — 7:16 — page 377 — #48

PROBLEMS

377

6.17 A series RLC circuit exhibits the following voltage and current responses: υC (t) = (6 cos 4t − 3 sin 4t)e

−2t

u(t) V,

If R = 12 , determine the values of Vs , L, and C. *6.22 Determine iC (t) in the circuit of Fig. 6.22 and plot its waveform for t ≥ 0.

iC (t) = −(0.24 cos 4t + 0.18 sin 4t)e−2t u(t) A. Determine α, ω0 , R, L, and C. *6.18

Determine iC (t) in the circuit of Fig. P6.18 for t ≥ 0.



8Ω iC

t=0

+ 12 V _

2H

υC

0.1 F

4Ω 2Ω 2Ω

4H

+ _ 30 V

t=0

12 Ω

iC υC

0.64 F

Figure P6.18: Circuit for Problem 6.18.

6.19

Determine υC (t) in the circuit of Fig. 6.19 for t ≥ 0.

t=0 4 μF

4 mA

Figure P6.22: Circuit for Problems 6.22 and 6.23.

6.23 Repeat Problem 6.22, retaining the same values for all elements in the circuit except C. Choose the value of C so that the response of iC (t) is critically damped. 6.24 Determine iC (t) in the circuit of Fig. 6.24 and plot its waveform for t ≥ 0, given that L = 0.05 H. Use a time scale that appropriately captures the shape of the waveform in your plot.

L

0.5 kΩ

0.52 Ω iC

υC

0.25 H

4 mA

0.1 Ω

t=0 0.1 Ω

1 F 1.8

υC

Figure P6.19: Circuit for Problem 6.19. Figure P6.24: Circuit for Problem 6.24 and 6.25.

6.20

Determine iC (t) in the circuit of Fig. 6.20 for t ≥ 0.



8Ω iC t=0

0.25 H



2.5 mF

*6.25 Choose the value of the inductor in the circuit of Fig. 6.24 so that υC exhibits a critically damped response and determine υC (t) for t ≥ 0. 6.26 Determine iC (t) in the circuit of Fig. 6.26 and plot its waveform for t ≥ 0, given that Vs = 24 V, R1 = 2 , R2 = 4 , L = 0.4 H, and C = 10 24 F.

+ 20 V _

R1

L iC

Figure P6.20: Circuit for Problem 6.20.

+ Vs _

t=0

C

R2 6.21 The circuit in Fig. 6-4(c) exhibits the response υ(t) = (12 + 36t)e−3t V,

(for t ≥ 0).

Figure P6.26: Circuit for Problems 6.26 and 6.27.

“book” — 2015/5/4 — 7:16 — page 378 — #49

378

CHAPTER 6

6.27 Repeat Problem 6.26 with the elements retaining their values, except change C to 10 29 F. 6.28

In the circuit of Fig. 6.28:

+ 0.2 V _

*(a) What is the value of υC (∞)? (b) How long does it take after t = 0 for υC to reach 0.99 of its final value? [Hint: After solving for υC (t), step through values of t over the range 2 ≤ t ≤ 2.5 to determine the value that satisfies the stated condition.]

iL

0.3 Ω

t=0

0.2 F

_ 0.1 V + Figure P6.30: Circuit for Problem 6.30.

6.31 Determine iC (t) and iL (t) in the circuit of Fig. 6.31 for t ≥ 0.

3A t=0 6Ω

0.2 H

0.1 Ω

RLC CIRCUITS

4 3

1H

Ω



+ 24 V _

10 Ω 0.25 F



100 Ω

iL

υC

iC

1H

t=0

+ 12 V _

Figure P6.28: Circuit for Problem 6.28.

5 mF

Figure P6.31: Circuit for Problem 6.31.

*6.29 Choose the value of C in the circuit of Fig. 6.29 so that υC (t) has a critically damped response for t ≥ 0. Plot the waveform of υC (t).

t=0

*6.32 For the circuit in Fig. P6.32, assume that before t = 0, the circuit had been in that state for a long time. Find υC (t) and iL (t) for t ≥ 0.

6Ω 6Ω

6Ω + 18 V _

12 Ω 0.1 H

C

υC

+ 12 V _

4V

+ _

+ υC _ 1 mH 6 μF

iL 2 mH

t=0 5Ω

Figure P6.29: Circuit for Problem 6.29. Figure P6.32: Circuit for Problem 6.32.

6.30 Determine iL (t) in the circuit of Fig. 6.30 and plot its waveform for t ≥ 0.

6.33

Find υC (t) for t ≥ 0 in the circuit in Fig. P6.33.

2A

“book” — 2015/5/4 — 7:16 — page 379 — #50

PROBLEMS

379 Section 6-7: Parallel RLC Circuit

2Ω 2Ω

10 V

+ _

4u(−t) V υC _ + + _



6.36 Determine iL (t) and iC (t) in the circuit of Fig. 6.36 and plot both waveforms for t ≥ 0. The SPDT switch was moved from position 1 to position 2 at t = 0.

8 mF 2Ω



2 mH 1

10 Ω

iC

10 6

H

Figure P6.36: Circuit for Problem 6.36.

6.37 Determine iL (t) in the circuit of Fig. 6.37 and plot its waveform for t ≥ 0.

10 Ω

t=0

3Ω + _ 24 V

t=0



0.1 F

6.34 For the circuit in Fig. P6.34, determine: (a) υC (0). (b) α, ω0 , and the type of response you expect υC (t) to exhibit. (c) iC (t) for t ≥ 0.

0.5 A

iL

t=0

+ 12 V _

Figure P6.33: Circuit for Problem 6.33.

2

12 Ω

iL

t=0 2 mH

iC 0.5 mF



6A

iC

4Ω 10 μF

8Ω 0.5 mH

+ _

Figure P6.37: Circuit for Problems 6.37 and 6.39.

+ _υC

*6.38 Determine iL (t) in the circuit of Fig. 6.38 and plot its waveform for t ≥ 0. The capacitor had no charge on it prior to t = 0.

6V

Figure P6.34: Circuit for Problem 6.34.

*6.35

For the circuit in Fig. P6.35, find υC (t) for t ≥ 0.



t=0

+ 1.5 mA _

iL 1.6 H

t=0 2 kΩ

0.1 μF

υC

0.5 mH 2A

8V + _

υs

+ υC _

Figure P6.38: Circuit for Problem 6.38.

4A 2 mF





Figure P6.35: Circuit for Problem 6.35.

6.39

Determine iC (t) in the circuit of Fig. 6.37 for t ≥ 0.

*6.40 Determine iL (t) in the circuit of Fig. 6.40 and plot its waveform for t ≥ 0.

“book” — 2015/5/4 — 7:16 — page 380 — #51

380

CHAPTER 6

500 Ω

RLC CIRCUITS

t=0 iC

+ 16 V _

iL

t=0

iL

2.5 μF

2.5 H



6Ω C

5 mH Figure P6.40: Circuit for Problems 6.40 and 6.41.

9V

6.41 Determine iC (t) in the circuit of Fig. 6.40 and plot its waveform for t ≥ 0.



+ _

Figure P6.44: Circuit for Problem 6.44.

6.42 Determine iL (t) in the circuit of Fig. 6.42 and plot its waveform for t ≥ 0. 6.45 25 3

200 Ω + 16 V _

H

0.2 mF

For the circuit in Fig. P6.45:

(a) Determine υC (t) for t ≥ 0.

iL t=0

800 Ω

(b) Determine the time at which the inductor has maximum energy stored in it and calculate the amount of that maximum energy.

Figure P6.42: Circuit for Problem 6.42.

*6.43 For the circuit of Fig. 6.43, determine: (a) iL (t) for t ≥ 0 (b) The amount of energy stored in the capacitor at t = ∞.



6 mH

0.5 mF



+ _υC

_ 12 V +

Figure P6.45: Circuit for Problem 6.45.

5Ω iL

+ _



6A

t=0

t=0

t=0



1Ω 10 V _ +



15 V

In the circuit in Fig. P6.46, υs = 20 V.

(a) Determine iL (t) for t ≥ 0.

1 mF 1 mH

5V

*6.46

+ _

(b) If the source is changed to υs (t) = e−2t u(t), can you still use the solution method in part (a) to find iL (t)? If not, why not?

Figure P6.43: Circuit for Problem 6.43.

6.44 Assume that the circuit in Fig. P6.44 had been in that state for a long time prior to t = 0. (a) Determine the value of C for which iL (t) exhibits the fastest smooth response. (b) Use the value of C found in part (a) to find iL (t) for t ≥ 0.

t=0

10 Ω υs

+ _

iL 10 Ω

2 mH

1 mF

+ _υC

Figure P6.46: Circuit for Problem 6.46.



“book” — 2015/5/4 — 7:16 — page 381 — #52

PROBLEMS

381

Section 6-8: General Solution

*6.50 The voltage in a certain circuit is described by the differential equation

6.47 The switch in the circuit of Fig. P6.47 was closed at t = 0 and then reopened at t = 1 ms. Determine iL (t) and υC (t) for t ≥ 0. Assume the capacitor had no charge prior to t = 0.

t = 1 ms iL + 3 mA _

1 kΩ

υ  + 5υ  + 6υ = 144

Determine υ(t) for t ≥ 0 given that υ(0) = 16 V and υ  (0) = 9.6 V/s. 6.51 The current in a certain circuit is described by the differential equation

t=0

3.2 H

(for t ≥ 0).

i  +

υC

0.2 μF



24 i  + 6i = 18

(for t ≥ 0).

Determine √ i(t) for t ≥ 0 given that i(0) = −2 A and i  (0) = 8 6 A/s. 6.52

Figure P6.47: Circuit for Problem 6.47.

For the circuit in Fig. P6.52:

(a) Determine iL (0) and υL (0). (b) Derive the differential equation for iL (t) for t ≥ 0. *6.48 After closing the switch in the circuit of Fig. P6.48 at t = 0, it was reopened at t = 1 ms. Determine iC (t) and plot its waveform for t ≥ 0. Assume no energy was stored in either L or C prior to t = 0.

200 Ω

*(c) Solve the differential equation and obtain an explicit expression for iL (t), given that Vs = 12 V, Rs = 3 , R1 = 0.5 , R2 = 1 , L = 2 H, and C = 2 F.

Rs

t = 1 ms

+ 20 V _

+ Vs _

iC

t=0

2

t=0 R1

2.5 μF

2.5 H

iL

L

1

R2

C

Figure P6.52: Circuit for Problem 6.52. Figure P6.48: Circuit for Problem 6.48.

6.49 Determine the current responses iL (t) and iC (t) to a rectangular-current pulse as shown in Fig. P6.49, given that Is = 10 mA and R = 499.99 . Plot the waveforms of iL (t), iC (t), and is (t) on the same scale.

6.53 Develop a differential equation for iL (t) in the circuit of Fig. P6.53. Solve it to determine iL (t) for t ≥ 0 subject to the following element values: Is = 36 μA, Rs = 100 k, R = 100 , L = 10 mH, and C = 10 μF.

R is =

Is 0

R

t=0

t = 1 ms

iL

iC

1H

1 μF

Figure P6.49: Circuit for Problem 6.49.

Ù Is ¥

iL

t=0 Rs

C

Figure P6.53: Circuit for Problem 6.53.

L

“book” — 2015/5/4 — 7:16 — page 382 — #53

382

CHAPTER 6

*6.54 Develop a differential equation for υC in the circuit of Fig. P6.54. Solve it to determine υC (t) for t ≥ 0. The element values are Is = 0.2 A, Rs = 30 , R1 = 10 , R2 = 20 , R3 = 20 , L = 4 H, and C = 5 mF.

t=0

R3

R1 Is

6.57 Repeat Problem 6.56, but this time assume that the switch had been closed for a long time and then opened at t = 0. *6.58 The op-amp circuit shown in Fig. P6.58 is called a multiple-feedback bandpass filter. If υin = A u(t), determine υout (t) for t ≥ 0 for A = 6 V, R1 = 10 k, R2 = 5 k, Rf = 50 k, and C1 = C2 = 1 μF.

L

Rs

C

C2

υC

Rf

R2 υin

iL

Figure P6.58: Circuit for Problem 6.58.

R1

R2

L

C

C2 υin

R1

R2

+ _

Figure P6.55: Circuit for Problem 6.55.

C1

Vs

+ _

R C1

R3

Figure P6.59: Circuit for Problem 6.59.

i2 t=0

υout

R4

*6.56 Determine i2 in the circuit of Fig. P6.56 for t ≥ 0, given that Vs = 10 V, Rs = 0.1 M, R = 1 M, C1 = 1 μF, and C2 = 2 μF.

Rs

υout

+

6.59 The op-amp circuit shown in Fig. P6.59 is called a two-pole low-pass filter. If υin = A u(t), determine υout (t) for t ≥ 0 for A = 2 V, R1 = 5 k, R2 = 10 k, R3 = 12 k, R4 = 20 k, C1 = 100 μF, and C2 = 200 μF.

t = 0.5 s

+ Vs _

_

R2

6.55 Develop a differential equation for iL in the circuit of Fig. P6.55. Solve it for t ≥ 0. The switch was closed at t = 0 and then reopened at t = 0.5 s, and the element values are Vs = 18 V, Rs = 1 , R1 = 5 , R2 = 2 , L = 2 H, and 1 F. C = 17

t=0

C1

R1

Figure P6.54: Circuit for Problem 6.54.

Rs

RLC CIRCUITS

C2

Figure P6.56: Circuit for Problems 6.56 and 6.57.

Section 6-9: Multisim 6.60 Using Multisim, draw a series RLC circuit with Vs = 24 V, R = 12 , L = 300 mH, and C = 10 mF. Use the Transient Analysis tool to obtain a plot of υC (t) for 0 < t < 0.2 s.

“book” — 2015/5/4 — 7:16 — page 383 — #54

PROBLEMS

383

R1

L1

90 Ω 500 mH + V1 _ 5u(0.003 − t) V R2

220 Ω

+ _

υC C1

100 μF

I1

0.1u(t − 0.05) A

Figure P6.64: Circuit for Problem 6.64.

6.61 Using Multisim, draw the circuit in Fig. E6.4 of Exercise 6.4. Use the Transient Analysis tool to obtain a plot of iC (t) for 0 < t < 1 s. 6.62 Using Multisim, draw the circuit in Fig. E6.4 of Exercise 6.4. Use the Transient Analysis tool to obtain three plots of iC (t) for (a) an underdamped response, (b) a critically damped response, and (c) an overdamped response. To obtain the three desired responses, adjust the value of the 20  resistor as needed. 6.63 Adjust the values of the source and the components in Fig. 6-23 such that the RLC circuit is excited and oscillates at a frequency of 1 MHz and the oscillation envelope decays to 10 percent of its initial value after 12 oscillations once the circuit is switched to “listen” mode. 6.64 Build the circuit shown in Fig. P6.64 in Multisim and then plot the voltage υC (t) from 0 to 200 ms using Transient Analysis. 6.65 Build the active second-order circuit shown in Fig. P6.65, plot the signal υout from 0 to 5 ms, and note how long it takes before the amplitude of the oscillations drops below 1 V. Change the value of R2 to 100 k and repeat the simulation. (You may need to readjust your timescale.) Potpourri Questions 6.66 How are transducers and actuators related? 6.67 How does a capacitive accelerometer work? 6.68 What are the differences between a passive RFID tag and an active RFID tag? 6.69 RFID tags operate at several frequency bands. How does the data speed change as the frequency is increased from the LF band to the microwave band? 6.70 Describe how electrical stimulation is used in a cochlear implant, in motor prostheses, and in reducing tremors in patients with Parkinson’s disease.

L1 10 mH C1 10 nF R2 R1 1 kΩ + V1 _ (1 + 4u(t − 0.001)) V

_

10 kΩ

+

υout

Figure P6.65: Circuit for Problem 6.65.

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m6.1 Initial and Final Conditions: The SPST switch in the circuit of Fig. m6.1 opens at t = 0, after it had been closed for a long time. Draw the circuit configurations that appropriately represent the state of the circuit at t = 0− , t = 0, and t = ∞ and use them to determine: (a) υC (0), iC (0) and υC (∞), and (b) iL (0), υL (0) and iL (∞). Component values are: R1 = 680 , R2 = 100 , R3 = 100 , switch resistance Rsw = 10 , wire resistance Rw = 10 , L = 3.3 mH, C = 0.1 μF, and Vs = 4.7 V.

“book” — 2015/5/4 — 7:16 — page 384 — #55

384

CHAPTER 6

R2

Rw

+

_ C

R3 Rsw R1

m6.3 Three-Resistor Circuit: Determine υ(t) of the circuit shown in Fig. m6.3 for t ≥ 0, given that the switch is opened at t = 0, after having been closed for a long time. Use the following component values: Vsrc = 8 V, R1 = 470 , R2 = 100 , Rw = 90 , C = 1.0 μF, and L = 33 mH. (a) Plot υ(t) from 0 to 5 ms using a tool such as MathScript or MATLAB. Include a hard copy of the script used to create the plot. (b) Determine the following values for υ(t):

L

υL(t)

iL(t)

iC(t)

+

_

υC(t)

+ _

t=0

RLC CIRCUITS

Vs

• Initial value υ(0), • Final value of υ(t), • Minimum value of υ(t), and • Time to reach the minimum value of υ(t).

Figure m6.1 Circuit for Problem m6.1.

t=0

m6.2 Natural Response of the Series RLC Circuit: The SPST switch in the circuit of Fig. m6.2 opens at t = 0, after it had been closed for a long time.

+ _

Vsrc

R1

R2

+ υ(t) _ C

(a) Determine υC (t) for t ≥ 0. (b) Plot υC (t) over the time range 0 ≤ t ≤ 1 ms with a plotting tool such as MathScript or MATLAB. (c) Determine the following numerical values; use either the equation υC (t) or take cursor measurements from the plot you created in the previous step: • Initial voltage υC , • υC (0), • Maximum value of υC , • Damped oscillation frequency fd = ωd /2π in Hz, and • Damping coefficient α. Use these component values: R1 = 220 , L = 33 mH, C = 0.01 μF, and Vsrc = 3.0 V.

R1 Vsrc

+ _

t=0 C

R2 = 330 ,

R2

+ _υC(t)

Figure m6.2 Circuit for Problem m6.2.

L

Rw

Figure m6.3 Circuit for Problem m6.3.

L

“book” — 2015/5/4 — 7:17 — page 385 — #1

7

CHAPTER

ac Analysis Contents 7-1 7-2 TB18 7-3 7-4 7-5 7-6 7-7 7-8 7-9 TB19 7-10 7-11 7-12 7-13

Overview, 386 Sinusoidal Signals, 386 Review of Complex Algebra, 389 Touchscreens and Active Digitizers, 393 Phasor Domain, 396 Phasor-Domain Analysis, 400 Impedance Transformations, 403 Equivalent Circuits, 410 Phasor Diagrams, 413 Phase-Shift Circuits, 416 Phasor-Domain Analysis Techniques, 420 Crystal Oscillators, 423 ac Op-Amp Circuits, 429 Op-Amp Phase Shifter, 431 Application Note: Power-Supply Circuits, 432 Multisim Analysis of ac Circuits, 437 Summary, 443 Problems, 444

Objectives

υ(t) Vm 0 T/2

T

3T/2

2T

t

−Vm Electric circuits whose currents and voltages vary sinusoidally with time—called alternating current (ac) circuits—are at the heart of most analog applications. This chapter and the next four are dedicated to ac circuits.

Learn to: 

Transform time-varying sinusoidal functions to the phasor domain and vice versa.



Analyze any linear circuit in the phasor domain.



Determine the impedance of any passive element, or the combination of elements connected in series or in parallel.



Perform source transformations, current division and voltage division, and determine Th´evenin and Norton equivalent circuits, all in the phasor domain.



Apply nodal analysis, mesh analysis, and other analysis techniques, all in the phasor domain.



Design simple RC phase-shift circuits.



Design a dc power-supply circuit.



Use Multisim to analyze ac circuits

“book” — 2015/5/4 — 7:17 — page 386 — #2

386

Overview From solar illumination to radio and cell-phone transmissions, we are surrounded by electromagnetic (EM) waves all of the time. EM waves are composed of sinusoidally varying electric and magnetic fields, and the fundamental parameter that distinguishes one EM wave from another is the wave’s frequency f (or equivalently, its wavelength λ = c/f , where c = 3 × 108 m/s is the velocity of light in a vacuum). The frequency of red light, for example, is 4.3 × 1014 Hz, and one of the frequencies assigned to cell-phone traffic is 1,900 MHz (1.9×109 Hz). Both are EM waves—and so are X-rays, infrared waves, and microwaves—but they oscillate sinusoidally at different frequencies and interact with matter differently (see Technology Brief 20 on the Electromagnetic Spectrum). The term “ac” (alternating current) is associated with electric circuits whose currents and voltages vary sinusoidally with time, just like EM waves. In fact, ac circuits and EM waves are not only similar, but they also are connected directly: when flowing in a conductor, an ac current with an oscillation frequency f radiates EM waves of the same frequency. The radiated waves can couple signals from one part of the circuit to another through the air space they share or the insulating regions between them. The coupling may serve as an intentional means of communication, as in the case of radio frequency identification (RFID) circuits, or it may introduce unwelcome signals that interfere with the intended operation of the circuit. Mitigation of such undesirable consequences is part of a subdiscipline of electrical engineering called electromagnetic compatibility. This and the next four chapters will be devoted to the study of ac circuits, which are far more prevalent than dc circuits and offer a much broader array of practical applications. In our study, we will assume that all currents and voltages are confined to the discrete elements in the circuit and to the connections between them, allowing us to ignore EM-compatibility issues altogether. In Chapter 12, we will learn how to use the Laplace transform technique to determine the response of a circuit to any source with any realistic waveform, including ac sources. In general, the solution consists of two components, a transient component—in response to sudden changes, such as the opening or closing of switches—and a steady state component that mimics the time variation of the source. If (a) all the sources in the circuit are ac sources and (b) our interest is in only the steady state component (because the transient component decays to approximately zero within a short time after connecting the circuit to the ac source), we can use the phasor domain technique (instead of the Laplace transform technique) to analyze the circuit, because it is mathematically

CHAPTER 7 AC ANALYSIS simpler and easier to implement. In fact, the phasor domain technique is a special case of the Laplace transform technique.  The phasor domain technique—also known as the frequency domain technique—applies to ac circuits only, and provides a solution of only the steady state component of the total solution. 

7-1

Sinusoidal Signals

The voltage between two points in a circuit (or the current flowing through a branch) is said to have a sinusoidal waveform if its time variation is given by a sinusoidal function. The term sinusoid includes both sine and cosine functions. For example, the expression υ(t) = Vm cos ωt

(7.1)

describes a sinusoidal voltage υ(t) that has an amplitude Vm and an angular frequency ω. The amplitude defines the maximum or peak value that υ(t) can reach, and −Vm is its lowest negative value. The argument of the cosine function, ωt, is measured either in degrees or in radians, with π (rad) ≈ 3.1416 (rad) = 180◦ .

(7.2)

Since ωt is measured in radians, the unit for ω is (rad/s). Figure 7-1(a) displays a plot of υ(t) as a function of ωt. The familiar cosine function starts at its maximum value (at ωt = 0), decreases to zero at ωt = π/2, goes into negative territory for half of a cycle, and completes its first cycle at ωt = 2π . Occasionally, we may want to display a sinusoidal signal as a function of t, instead of ωt. We note that the angular frequency ω is related to the oscillation frequency (or simply the frequency) f of the signal by ω = 2πf

(rad/s),

(7.3)

with f measured in hertz (Hz), which is equivalent to cycles/second. A sinusoidal voltage with a frequency of 100 Hz makes 100 oscillations in 1 s, each of duration 1/100 = 0.01 s. The duration of a cycle is its period T . Thus, T =

1 f

(s).

(7.4)

By combining Eqs. (7.1), (7.3), and (7.4), υ(t) can be rewritten as 2π t , (7.5) υ(t) = Vm cos T

“book” — 2015/5/4 — 7:17 — page 387 — #3

7-1

SINUSOIDAL SIGNALS

387

Table 7-1: Useful trigonometric identities (additional

υ(t)

relations are given in Appendix D).

Vm 0

π/2

π

−Vm







ωt

υ(t) versus ωt

(a) υ(t) Vm 0 T/2 −Vm

T

3T/2

2T

sin x = ± cos(x ∓ 90◦ ) cos x = ± sin(x ± 90◦ ) sin x = − sin(x ± 180◦ ) cos x = − cos(x ± 180◦ ) sin(−x) = − sin x cos(−x) = cos x

(7.7a) (7.7b) (7.7c) (7.7d) (7.7e) (7.7f)

sin(x ± y) = sin x cos y ± cos x sin y cos(x ± y) = cos x cos y ∓ sin x sin y

(7.7g) (7.7h)

2 sin x sin y = cos(x − y) − cos(x + y) 2 sin x cos y = sin(x + y) + sin(x − y) 2 cos x cos y = cos(x + y) + cos(x − y)

(7.7i) (7.7j) (7.7k)

t In addition to ωt, the argument of the cosine function contains a constant angle of −60◦ . A cosine-referenced sinusoidal function generally takes the form

υ(t) versus t

(b)

υ(t) = Vm cos(ωt + φ),

Figure 7-1: The function υ(t) = Vm cos ωt plotted as a function of (a) ωt and (b) t.

which is displayed in Fig. 7-1(b) as a function of t. We observe that the cyclical pattern of the waveform repeats itself every T seconds. That is, υ(t) = υ(t + nT ) (7.6) for any integer value of n. Sinusoidal waveforms can be expressed in terms of either sine or cosine functions.  To avoid confusion, we adopt the cosine form as our reference standard throughout this and followup chapters.  This means that we will always express voltages and currents in terms of cosine functions, so if a voltage (or current) waveform is given in terms of a sine function, we should first convert it to a cosine form with a positive amplitude before proceeding with our circuit analysis. Conversion from sine to cosine form is realized through the application of Eq. (7.7a) of Table 7-1. For example, i(t) = 6 sin(ωt + 30◦ ) = 6 cos(ωt + 30◦ − 90◦ ) = 6 cos(ωt − 60◦ ).

(7.8)

(7.9)

where φ is called its phase angle. For i(t) of Eq. (7.8), φ = −60◦ . The angle φ may assume any positive or negative value, but we usually add or subtract multiples of 2π radians (or equivalently, multiples of 360◦ ) so that the remainder is between −180◦ and +180◦ . The magnitude and sign (+ or −) of φ determine, respectively, by how much and in what direction the waveform of υ(t) is shifted along the time axis, relative to the reference waveform corresponding to υ(t) with φ = 0. Figure 7-2 displays three waveforms:  υ1 (t) = Vm cos

2π t π − T 4

 (lags by π/4),

(7.10a)

2π t (reference waveform with φ = 0), T (7.10b)   2π t π υ3 (t) = Vm cos + (leads by π/4). (7.10c) T 4

υ2 (t) = Vm cos

We observe that waveform υ3 (t), which is shifted backwards in time relative to the reference waveform υ2 (t), attains its peak value before υ2 (t) does. Consequently, waveform υ3 (t) is said to lead υ2 (t) by a phase lead of π/4. Similarly, waveform υ1 (t) lags υ2 (t) by a phase lag of π/4. A cosine function with a negative phase angle φ takes longer to reach a specified

“book” — 2015/5/4 — 7:17 — page 388 — #4

388

CHAPTER 7 AC ANALYSIS

υ

υ3(t): Leads reference wave (occurs earlier in time)

Vm

υ2(t): Reference wave (ϕ = 0) υ1(t): Lags reference wave (occurs later in time)

ϕ = π/4

ϕ = −π/4

T 2

T

3T 2

t

∆t −Vm Figure 7-2: Plots of υ(t) = Vm cos[(2πt/T ) + φ] for three different values of φ.

reference level (such as the peak value) than it takes the zerophase angle function to reach that level, signifying a phase lag. When φ is positive, it signifies a phase lead. A phase angle of 2π corresponds to a time shift along the time axis equal to one full period T . Proportionately, a phase angle of φ (in radians) corresponds to a time shift t given by   φ T. (7.11) t = 2π We generalize our discussion of phase lead and lag by stating that:  Given two sinusoidal functions with the same angular frequency ω, and both expressed in standard cosine form as υ1 (t) = V1 cos(ω + φ1 ) and

Example 7-1: Voltage Waveform

A sampling oscilloscope is used to measure a voltage signal υ(t). The measurements reveal that υ(t) is periodic with an amplitude of 10 V, its maxima are separated by 20 ms, and one of its maxima occurs at t = 1.2 ms. Determine the functional form of υ(t). Solution: Given that Vm = 10 V and T = 20 ms = 2 × 10−2 s, υ(t) is given by 

2π t +φ υ(t) = 10 cos 2 × 10−2

 = 10 cos(100π t + φ) V.

Application of υ(t = 1.2 ms) = 10 V gives 10 = 10 cos(100π × 1.2 × 10−3 + φ),

υ2 (t) = V2 cos(ω + φ2 ),

which requires the argument of the cosine to be a multiple of 2π ,

the relevant terminology is: υ2 leads υ1 by (φ2 − φ1 ),

0.12π + φ = 2nπ,

υ2 lags υ1 by (φ1 − φ2 ), υ1 and υ2 are in phase

if φ2 = φ1 ,

υ1 and υ2 are in phase-opposition

if φ2 = φ1 ± 180◦ .

n = 0, ±1, ±2, . . .

The smallest value of φ in the range [−180◦ , 180◦ ] that satisfies the preceding equation corresponds to n = 0, and is given by φ = −0.12π = −21.6◦ .

(out of phase) 

Hence, υ(t) = 10 cos(100π t − 21.6◦ ) V.

“book” — 2015/5/4 — 7:17 — page 389 — #5

7-2

REVIEW OF COMPLEX ALGEBRA

389

Example 7-2: Phase Lead / Lag

Exercise 7-2: Given two current waveforms:

i1 (t) = 3 cos ωt

Given the current waveforms and

i1 (t) = −8 cos(ωt − 30◦ ) A

i2 (t) = 3 sin(ωt + 36◦ ),

and

does i2(t) lead or lag i1(t), and by what phase angle?

i2 (t) = 12 sin(ωt + 45◦ ) A,

Answer: i2(t) lags i1(t) by 54◦. (See

)

does i1 (t) lead i2 (t), or the other way around, and by how much? Solution: Standard cosine format requires that the sinusoidal functions be cosines and that the amplitudes have positive values. Application of Eq. (7.7d) of Table 7-1 allows us to remove the negative sign preceding the amplitude of i1 (t), i1 (t) = −8 cos(ωt − 30◦ ) = 8 cos(ωt − 30◦ + 180◦ ) = 8 cos(ωt + 150◦ ) A. Application of Eq. (7.7a) to i2 (t) leads to

7-2

Review of Complex Algebra

This section provides a review of complex algebra, in preparation for the introduction of the phasor domain technique in Section 7-3. A complex number z may be written in the rectangular form z = x + jy,

i2 (t) = 12 sin(ωt + 45◦ ) = 12 cos(ωt + 45◦ − 90◦ ) = 12 cos(ωt − 45◦ ) A. Hence, φ1 = 150◦ , φ2 = −45◦ , and

(7.12)

where x and y are the real √ (Re) and imaginary (Im) parts of z, respectively, and j = −1. That is, x = Re(z),

y = Im(z).

(7.13)



φ = φ2 − φ1 = −195 . Alternatively, z may be written in polar form as The concept of phase lead/lag requires that φ be within the range [−180◦ , 180◦ ]. Addition of 360◦ to φ converts it to 165◦ , which means that i2 leads i1 by 165◦ . Concept Question 7-1: A sinusoidal waveform is

characterized by three parameters. What are they, and what does each one of them specify? (See )

z = |z|ej θ = |z| θ

(7.14)

where |z| is the magnitude of z, θ is its phase angle, and the form θ is a useful shorthand representation commonly used in numerical calculations. A phase angle may be expressed in degrees, as in θ = 30◦ , or in radians, as in θ = 0.52 rad. By applying Euler’s identity,

Concept Question 7-2: Waveforms υ1 (t) and υ2 (t) have

the same angular frequency, but υ1 (t) leads υ2 (t). Will the peak value of υ1(t) occur sooner or later than that of υ2(t)? Explain. (See )

Exercise 7-1: Provide an expression for a 100 V, 60 Hz

ej θ = cos θ + j sin θ,

(7.15)

we can convert z from polar form, as in Eq. (7.14), into rectangular form, as in Eq. (7.12)),

voltage that exhibits a minimum at t = 0. Answer: υ(t) = 100 cos(120πt + 180◦) V. (See

)

z = |z|ej θ = |z| cos θ + j |z| sin θ,

(7.16)

“book” — 2015/5/4 — 7:17 — page 390 — #6

390

CHAPTER 7 AC ANALYSIS

Im(z)

x = |z| cos θ y = |z| sin θ

z

y

both cases. Also note that, since |z| is a positive quantity, only the positive root in Eq. (7.18) is applicable. The complex conjugate of z, denoted with a star superscript (or asterisk), is obtained by replacing j (wherever it appears) with −j , so that

+

|z| = x2 + y2 θ = tan−1 (y/x)

|z|

z∗ = (x + jy)∗ = x − jy = |z|e−j θ = |z| −θ .

θ

Re(z)

x

The magnitude |z| is equal to the positive square root of the product of z and its complex conjugate:

Figure 7-3: Relation between rectangular and polar representations of a complex number z = x + jy = |z|ej θ .

|z| =

√ z z∗ .

(7.20)

We now highlight some of the properties of complex algebra that we will likely encounter in future sections.

which leads to the relations x = |z| cos θ,  |z| = x 2 + y 2 ,

(7.19)

y = |z| sin θ,

(7.17)

θ = tan−1 (y/x).

(7.18)

The two forms of z are illustrated graphically in Fig. 7-3. Because in the complex plane, a complex number assumes the form of a vector, it is represented by a bold letter in this book. When using Eq. (7.18), care should be taken to ensure that θ is in the proper quadrant by noting the signs of x and y individually, as illustrated in Fig. 7-4. Complex numbers z2 and z4 point in opposite directions and their phase angles θ2 and θ4 differ by 180◦ , despite the fact that (y/x) has the same value in

Equality: If two complex numbers z1 and z2 are given by z1 = x1 + jy1 = |z1 |ej θ1 ,

(7.21a)

z2 = x2 + jy2 = |z2 |ej θ2 ,

(7.21b)

then z1 = z2 if and only if (iff) x1 = x2 and y1 = y2 or, equivalently, |z1 | = |z2 | and θ1 = θ2 . Addition: z1 + z2 = (x1 + x2 ) + j (y1 + y2 ).

(7.22)

Multiplication: z1 z2 = (x1 + jy1 )(x2 + jy2 )

Im(z) z2 = −2 + j3

θ2 = 180o − θ1

3 θ2 2 1

= (x1 x2 − y1 y2 ) + j (x1 y2 + x2 y1 ),

z1 = 2 + j3

θ1

θ1 =

3 tan−1 2

2 3 −3 −2 −1 1 −1 θ3 = −θ2 θ θ4 θ4 = −θ1 −2 3 z3 = −2 − j3

−3

(7.23a)

or

= 56.3o Re(z)

z4 = 2 − j3

Figure 7-4: Complex numbers z1 to z4 have the same  magnitude |z| = 22 + 32 = 3.61, but their polar angles depend on the polarities of their real and imaginary components.

z1 z2 = |z1 |ej θ1 · |z2 |ej θ2 = |z1 ||z2 |ej (θ1 +θ2 ) = |z1 ||z2 |[cos(θ1 + θ2 ) + j sin(θ1 + θ2 )].

(7.23b)

Division: For z2  = 0, z1 x1 + jy1 (x1 + jy1 ) (x2 − jy2 ) = = · z2 x2 + jy2 (x2 + jy2 ) (x2 − jy2 ) (x1 x2 + y1 y2 ) + j (x2 y1 − x1 y2 ) = , x22 + y22 (7.24a)

“book” — 2015/5/4 — 7:17 — page 391 — #7

7-2

REVIEW OF COMPLEX ALGEBRA

391

Table 7-2: Properties of complex numbers. Euler’s Identity: ej θ = cos θ + j sin θ sin θ =

ej θ − e−j θ 2j

z = x + jy = |z|ej θ

cos θ =

ej θ + e−j θ 2

z∗ = x − jy = |z|e−j θ

zn = |z|n ej nθ

 √ |z| = zz∗ = x 2 + y 2 ⎧ −1 tan (y/x) ⎪ ⎪ ⎪ ⎨tan−1 (y/x) ± π θ= ⎪ π/2 ⎪ ⎪ ⎩ −π/2 z1/2 = ±|z|1/2 ej θ/2

z1 = x1 + jy1

z2 = x2 + jy2

z1 = z2 iff x1 = x2 and y1 = y2

z1 + z2 = (x1 + x2 ) + j (y1 + y2 )

z1 z2 = |z1 ||z2 |ej (θ1 +θ2 )

|z1 | j (θ1 −θ2 ) z1 = e z2 |z2 |

x = Re(z) = |z| cos θ

y = Im(z) = |z| sin θ

if x if x if x if x

> 0, < 0, = 0 and y > 0, = 0 and y < 0.

−1 = ej π = e−j π = 1 ±180◦ j = ej π/2 = 1 90◦  (1 + j ) j = ±ej π/4 = ± √ 2

−j = e−j π/2 = 1 −90◦  (1 − j ) −j = ±e−j π/4 = ± √ 2

or

Useful Relations: |z1 |ej θ1 |z1 | j (θ1 −θ2 ) z1 = = e j θ 2 z2 |z2 |e |z2 | |z1 | = [cos(θ1 − θ2 ) + j sin(θ1 − θ2 )]. |z2 | (7.24b)

Powers: For any positive integer n,

−1 = ej π = e−j π = 1 180◦ , j =e

j π/2

= 1 90◦ ,

−j = −ej π/2 = e−j π/2 = 1 −90◦ ,  ±(1 + j ) j = (ej π/2 )1/2 = ±ej π/4 = , √ 2  ±(1 − j ) −j = ±e−j π/4 = . √ 2

(7.27a) (7.27b) (7.27c) (7.27d) (7.27e)

zn = (|z|ej θ )n = |z|n ej nθ = |z|n (cos nθ + j sin nθ ),

(7.25)

z1/2 = ±|z|1/2 ej θ/2 = ±|z|1/2 [cos(θ/2) + j sin(θ/2)].

(7.26)

For quick reference, the preceding properties of complex numbers are summarized in Table 7-2. Note that if a complex number is given by (a + j b) and b = 1, it can be written either as (a + j 1) or simply as (a + j ). Thus, j is synonymous with j 1.

“book” — 2015/5/4 — 7:17 — page 392 — #8

392

CHAPTER 7 AC ANALYSIS Since our preference is to end up with a phase angle within the range between −180◦ and +180◦ , we will choose −180◦ . Hence,

Im



−3

−2 θI

|I|

I = 3.61e−j 123.7 .

Re

θV

(b)

|V|

VI = (5 −53.1◦ )(3.61 −123.7◦ )

−3

I

= (5 × 3.61) (−53.1◦ − 123.7◦ ) = 18.05 −176.8◦ .

−4

V

Figure 7-5: Complex numbers V and I in the complex plane

(c)

(Example 7-3).





V 5e−j 53.1 j 70.6◦ . = ◦ = 1.39e −j 123.7 I 3.61e

Given two complex numbers (e)

V = 3 − j 4,



I = −(2 + j 3).

Solution: (a)  √ VV∗ = (3 − j 4)(3 + j 4) = 9 + 16 = 5,

θV = tan−1 (−4/3) = −53.1◦ ,

 ◦ 3.61e−j 123.7 √ ◦ ◦ = ± 3.61 e−j 123.7 /2 = ±1.90e−j 61.85 .

I=

(a) Express V√and I in polar form, and find (b) VI, (c) VI∗ , (d) V/I, and (e) I .

|V| =



(d)

Example 7-3: Working with Complex Numbers





VI∗ = 5e−j 53.1 × 3.61ej 123.7 = 18.05ej 70.6 .

Concept Question 7-3: If Z is a complex number that lies in the first quadrant in the complex plane, its complex conjugate Z∗ will lie in which quadrant? (See )



V = |V|ej θV = 5e−j 53.1 = 5 −53.1◦ ,  √ |I| = 22 + 32 = 13 = 3.61.

Concept Question 7-4: If two complex numbers have the

same magnitude, are they necessarily equal to each other? (See )

Since I = (−2 − j 3) is in the third quadrant in the complex plane (Fig. 7-5), θI = −180◦ + tan

−1 3 2

Exercise 7-3: Express the following complex functions

in polar form:

= −123.7◦ ,

I = 3.61 −123.7◦ .

z1 = (4 − j 3)2 ,

Alternatively, whenever the real part of a complex number is negative, we can factor out a (−1) multiplier and then use Eq. (7.27a) to replace it with a phase angle of either +180◦ or −180◦ , as needed. In the case of I, the process is as follows: ◦

I = −2 − j 3 = −(2 + j 3) = e±j 180 ·

 ◦

22 + 32 ej tan ◦

= 3.61ej 57.3 e±j 180 .

z2 = (4 − j 3)1/2 .

z1 = 25 −73.7◦ ,

Answer:

(See

√ z2 = ± 5 −18.4◦ .

)

−1 (3/2)

Exercise 7-4: Show that



2j = ±(1 + j ). (See

)

“book” — 2015/5/4 — 7:17 — page 393 — #9

TECHNOLOGY BRIEF 18: TOUCHSCREENS AND ACTIVE DIGITIZERS

Technology Brief 18 Touchscreens and Active Digitizers Touchscreen is the common name given to a wide variety of technologies that allow computer displays to directly sense information from the user. In older systems, this usually meant the display could detect and pinpoint where a user touched the screen surface; newer systems can detect multiple touch locations as well as the associated touch pressures simultaneously, with very high resolution. This has led to a surge of applications in mobile computing, cell phones, personal digital assistants (PDA), and consumer appliances. Interactive touchscreens which detect multiple touches and interact with styli are now commonly used in phones, tablet computers and e-readers. Numerous technologies have been developed since the invention of the electronic touch interface in 1971 by Samuel C. Hurst. Some of the earlier technologies were susceptible to dust, damage from repeat use, and poor transparency. These issues largely have been resolved over the years (even for older technologies) as experience and advanced material selection have led to improved devices. With the explosion of consumer interest in portable, interactive electronics, newer technologies have emerged that are more suitable for these applications. Figure TF18-1 summarizes the general categories of touchscreens in use today. Historically, touchscreens were manufactured separately from displays and added as an extra layer of the display. More recently, display companies have begun to manufacture sensing technology directly into the displays; some of the newer technologies reflect this.

Resistive Resistive touchscreens are perhaps the simplest to understand. A thin, flexible membrane is separated from a plastic base by insulating spacers. Both the thin membrane and the plastic base are coated on the inside with a transparent conductive film (indium tin oxide (ITO) often is used). When the membrane is touched, the two conductive surfaces come into contact. Detector circuits at the edges of the screen can detect this change in resistance between the two membranes and pinpoint the location on the X–Y plane. Older designs of this type were susceptible to membrane damage (from repeated flexing) and suffered from poor transparency.

393

Capacitive Older capacitive touchscreens employ a single thin, transparent conductive film (usually indium tin oxide (ITO)) on a plastic or glass base. The conductive film is coated with another thin, transparent insulator for protection. Since the human body stores charge, a finger tip moved close to the surface of the film effectively forms a capacitor where the film acts as one of the plates and the finger as the other. The protective coating and the air form the intervening dielectric insulator. This capacitive coupling changes how a current flowing across the film surface is distributed; by placing electrodes at the screen corners and applying an ac electric signal, the location of the finger capacitance can be calculated precisely. One variant of this idea is to divide the sensing area into many smaller squares (just like pixels on the display) and to sense the change in capacitance across each of them continuously and independently; this is commonly known as self-capacitance sensing. A newer development, found in many modern portable devices, is the use of mutual capacitance sensing touchscreens, which employ two sets of conductive lines, each on a different layer. On one layer, the lines might run horizontally, while on another layer below the first the lines run vertically. At each point of overlap between the lines on the two layers, a parallel plate capacitor is formed. If there are M lines on the top layer and N lines on the bottom, there will be M ×N such nodes.Whenever a finger moves near a node, the capacitance of the node changes. By monitoring the capacitance of each node continuously, the touchscreen can detect when touches occur and where. The principal advantages of a touchscreen of this type are its ability to detect many simultaneous touches and its ability to detect very light ones. Capacitive technologies are much more resistant to wear and tear (since they are not flexed) than resistive touchscreen and are somewhat more transparent (> 85 percent transparency) since they can have fewer films and avoid air gaps. These types of screens can be used to detect metal objects as well, so pens with conductive tips can be used on writing interfaces. Not all capacitive touch systems are integrated with screens; a number of interactive media technologies developed over the last 15 years integrate the touch sensing technology into furniture, household objects, or even countertops and overlay a display using nearby projection equipment. Some interactive tables operate this way. A completely different way to detect touch relies on the measurement of acoustic energy on or near the touchscreen. There are several ways to make use of

“book” — 2015/5/4 — 7:17 — page 394 — #10

394

TECHNOLOGY BRIEF 18: TOUCHSCREENS AND ACTIVE DIGITIZERS

Membrane Conductive film

Membrane

Conductive film Cfinger

Spacer Plastic

Plastic

(a) Resistive

(b) Capacitive

Acoustic emitter Strain sensor

Strain sensor

Force Stress

Acoustic sensor

Stress

5 MHz Acoustic wave

(d) Acoustic

LC R

(c) Pressure

Screen

Acoustic dampening

LED

IR beam

Pen Electromagnetic radiation

Detector

Screen Wires

(e) Infrared

(f) Active digitizer

Figure TF18-1: Touchscreen technologies: (a) resistive, (b) capacitive, (c) pressure/strain sensor, (d) acoustic, (e) infrared, and (f) active digitizer.

acoustic energy to measure touch. One implementation relies on transmission of high-frequency acoustic energy across the surface of the display material.

Pressure Touch also can be detected mechanically. Pressure sensors can be placed at the corners of the display screen

or even the entire display assembly, so whenever the screen is depressed, the four corners will experience different stresses depending on the (X,Y) position of the pressure point. Pressure screens benefit from high resistance to wear and tear and no losses in transparency (since there is no need to add layers over the display screen).

“book” — 2015/5/4 — 7:17 — page 395 — #11

TECHNOLOGY BRIEF 18: TOUCHSCREENS AND ACTIVE DIGITIZERS

Acoustic A completely different way to detect touch relies on the transmission of high-frequency acoustic energy across the surface of the display material. Bursts of 5 MHz tones are launched by acoustic actuators from two corners of the screen. Acoustic reflectors all along the edges of the screen re-direct the incoming waves to the sensors. Any time an object comes into contact with the screen, it dampens or absorbs some fraction of the energy traveling across the material. The exact (X,Y) position can be calculated from the energy hitting the acoustic sensors. The contact force can be calculated as well, because the acoustic energy is dampened more or less depending on how hard the screen is pressed. Another approach is to listen, with very sensitive acoustic transducers (i.e., microphones) to the characteristic pressure signal (e.g., sound) made in the touchscreen material when it is touched. By placing several transducers around the edge of the screen, the system can determine if a touch occurred and where. One drawback is that motionless fingers cannot be detected. However, this does provide an advantage in that resting objects (i.e., your cheek) do not trigger the screen. This method is sometimes known as acoustic pulse recognition.

Infrared One of the oldest and least used technologies is the infrared touchscreen. This technology relies on infrared emitters (usually infrared diodes) aligned along two adjoining edges of the screen and infrared detectors aligned across from the emitters at the other two edges. The position of a touch event can be determined through a process based on which light paths are interrupted. The detection of multiple simultaneous touch events is possible. Infrared screens are somewhat bulky, prone to damage or interference from dust and debris, and need special modifications to work in daylight.They largely have been displaced by newer technologies.

Electromagnetic Resonance Another technology in widespread use is the electromagnetic resonance detection scheme used by many tablet PCs. Strictly speaking, many tablet PC screens are not touchscreens; they are called active digitizers because they can detect the presence and location of the tablet pen as it approaches the screen (even without contact). In this scheme, a very thin wire grid is integrated within

395

the display screen (which usually is a flat-profile LCD display). The pen itself contains a simple RLC resonator (see Section 6-1) with no power supply. The wire grid alternates between two modes (transmit and receive) every ∼ 20 milliseconds. The grid essentially acts as an antenna. During the transmit mode, an ac signal is applied to the grid and part of that signal is emitted into the air around the display. As the pen approaches the grid, some energy from the grid travels across to the pen’s resonator which begins to oscillate. In receive mode, the grid is used to “listen” for ac signals at the resonator frequency; if those signals are present, the grid can pinpoint where they are across the screen. A tuning fork provides a good analogy. Imagine a surface vibrating at a musical note; if a tuning fork designed to vibrate at that note comes very close to that surface, it will begin to oscillate at the same frequency. Even if we were to stop the surface vibrations, the tuning fork will continue to make a sound for a little while longer (as the resonance dies down). In a similar way, the laptop screen continuously transmits a signal and listens for the pen’s electromagnetic resonance. Functions (such as buttons and pressure information) can be added to the pen by having the buttons change the capacitance value of the LCR when pressed; in this way, the resonance frequency will shift (see Section 6-2), and the shift can be detected by the grid and interpreted as a button press.

Increased Integration Mobile devices have largely driven the development of advanced touch technologies in the last few years. Given the constant pressure to miniaturize and integrate, a number of companies have or are developing integrated touch and display systems. Unlike the earlier-generation technologies, the display and the touch sensor are not manufactured separately and then integrated during assembly. Rather, the touch sensor conductors (in the case of capacitive sensing) are designed into the very display itself, either into the conductive traces in/on the pixels of the display or immediately over them. In other designs, light-sensing pixels are manufactured into each display pixel of a display, giving the display not only the ability to produce images but also to sense nearby objects that occlude light landing on the sensing pixels. Even the integrated circuits are increasingly being integrated; earlier-generation systems relied on stand-alone touch controller IC chips that managed the sensor information and communicated it to the application processor in the mobile devices. There is a push to integrate this functionality into some phone processors directly.

“book” — 2015/5/4 — 7:17 — page 396 — #12

396

CHAPTER 7 AC ANALYSIS

7-3 Phasor Domain

In general, the phasor-domain quantity X is complex, consisting of a magnitude |X| and a phase angle φ,

In this chapter, we explore how currents and voltages defined in the time domain are transformed into their counterparts in the phasor domain (also called the frequency domain), and why such a transformation facilitates the analysis of ac circuits. The KVL and KCL equations characterizing an ac circuit containing capacitors and inductors take the form of integrodifferential equations with forcing functions (representing the real sources in the circuit) that vary sinusoidally with time. The phasor technique allows us to transform the equations from the time domain to the phasor domain, as a result of which the integro-differential equations get converted into linear equations with no sinusoidal functions. After solving for the desired variable—such as a particular voltage or current— in the phasor domain, conversion back to the time domain provides the same solution that we would have obtained had we solved the integro-differential equations entirely in the time domain. The procedure involves multiple steps, but it avoids the complexity of solving differential equations containing sinusoidal functions.

X = |X|ej φ . Using this expression in Eq. (7.28) gives

x(t) = Re[|X|ej φ ej ωt ] = Re[|X|ej (ωt+φ) ] = |X| cos(ωt +φ). (7.30) Application of the Re operator allows us to transform a function from the phasor domain to the time domain. The reverse operation, namely to specify the phasor-domain equivalent of a time function, can be ascertained by comparing the two sides of Eq. (7.30). Thus, for a voltage υ(t) with phasor counterpart V, the correspondence between the two domains is as follows: Time Domain

Phasor Domain

υ(t) = V0 cos ωt

V = V0

υ(t) = V0 cos(ωt + φ)

V = V0 e .

(7.31b)

V = V0 e−j π/2 .

(7.32)

Since cos(ωt − π/2) = cos(π/2 − ωt) = sin ωt and e−j π/2 = cos(π/2) − j sin(π/2) = −j ,

Transformation from the time domain to the phasor domain entails transforming all time-dependent quantities in the circuit, which in effect transforms the entire circuit from the time domain to an equivalent circuit in the phasor domain. The quantities involved in the transformation include all currents and voltages, all sources, and all capacitors and inductors. The values of capacitors and inductors do not change per se, but their i–υ relationships undergo a transformation because they involve differentiation or integration with respect to t. Any cosinusoidally time-varying function x(t), representing a voltage or a current, can be expressed in the form x(t) = Re[  X ej ωt ],

(7.31a) jφ

If φ = −π/2, υ(t) = V0 cos(ωt − π/2)

7-3.1 Time-Domain/Phasor-Domain Correspondence

(7.29)

(7.28)

Eq. (7.32) reduces to υ(t) = V0 sin ωt

V = −j V0 ,

(7.33)

V = V0 ej (φ−π/2) .

(7.34)

which can be generalized to υ(t) = V0 sin(ωt + φ)

Occasionally, voltage and current time functions may encounter differentiation or integration. For example, consider a current i(t) with a corresponding phasor I, i(t) = Re[Iej ωt ],

(7.35)

where X is a time-independent function called the phasor counterpart of x(t). Thus, x(t) is defined in the time domain, while its counterpart X is defined in the phasor domain.

where I may be complex but, by definition, not a function of time. The derivative di/dt is given by   d d di j ωt j ωt = [Re(Ie )] = Re (Ie ) = Re[j ωI ej ωt ],  dt dt dt

 To distinguish phasor quantities from their timedomain counterparts, phasors are always represented by bold letters in this book. 

(7.36) where in the second step we interchanged the order of the two operators, Re and d/dt, which is justified by the fact that the two operators are independent of one another, meaning that taking the real part of a quantity has no influence on taking its

phasor

phasor of di/dt

“book” — 2015/5/4 — 7:17 — page 397 — #13

7-3

PHASOR DOMAIN

397

time derivative, and vice versa. We surmise from Eq. (7.36) that

Table 7-3: Time-domain sinusoidal functions x(t) and their cosine-reference phasor-domain counterparts X, where x(t) = Re [Xej ωt ].

di dt

j ωI,

(7.37) x(t)

X

A cos ωt

A

A cos(ωt + φ)

Aej φ

 Differentiation of a time function i(t) in the time domain is equivalent to multiplication of its phasor counterpart I by j ω in the phasor domain. 

−A cos(ωt + φ)

Aej (φ±π)

A sin ωt

Ae−j π/2 = −j A

A sin(ωt + φ)

Aej (φ−π/2)

Similarly,

−A sin(ωt + φ)

Aej (φ+π/2)

or:



 i dt =

Re[Ie 

j ωt

] dt

   I j ωt Iej ωt dt = Re , e jω 

= Re

phasor of

(7.38)

i dt

or  i dt

I , jω

j ωAej φ 1 X jω 1 Aej φ jω

iR = Re[IR ej ωt ].

(7.41b)

Inserting these expressions into Eq. (7.40) gives

 Integration of i(t) in the time domain is equivalent to dividing its phasor I by j ω in the phasor domain.  Table 7-3 provides a summary of some time functions and their phasor-domain counterparts.

Impedance of Circuit Elements

The υ–i relationship for a resistor R is υR = RiR .

(7.40)

If iR is a sinusoidal function of t, the same is true for υR . The time-domain quantities υR and iR are related to their phasordomain counterparts by υR = Re[VR ej ωt ]

Re[VR ej ωt ] = R Re[IR ej ωt ] = Re[RIR ej ωt ].

(7.42)

Upon combining both sides under the same real-part (Re) operator, we have Re[(VR − RIR )ej ωt ] = 0.

(7.43a)

Through a somewhat similar treatment that uses a sine reference—rather than a cosine reference—to define sinusoidal functions, we can obtain the result

Resistors

and

j ωX

(7.39)

which states that:

7-3.2

d (x(t)) dt d [A cos(ωt + φ)] dt  x(t) dt  A cos(ωt + φ) dt

(7.41a)

Im[(VR − RIR )ej ωt ] = 0,

(7.43b)

which, for the sake of expediency, we simply state without taking the steps to prove it. In view of Eqs. (7.43a) and (7.43b), both the real and imaginary components of the quantity inside the square bracket are zero. Hence, the quantity itself is zero, and since ej ωt  = 0, it follows that VR − RIR = 0.

(7.44)

“book” — 2015/5/4 — 7:17 — page 398 — #14

398

CHAPTER 7 AC ANALYSIS

In the phasor domain:

Capacitors

 The impedance Z of a circuit element is defined as the ratio of the phasor voltage across it to the phasor current entering through its plus (+) terminal, 

Z=

V I

( ),

(7.45)

and the unit of Z is the ohm ( ). For a resistor, Eq. (7.44) gives ZR =

VR = R. IR

(7.46)

Thus, for a resistor the impedance is entirely real, and the form of the υ–i relationship is the same in both the time and phasor domains.

Since for a capacitor iC = C

dυC , dt

(7.52)

it follows that in the phasor domain, IC = j ωCVC

(7.53)

and the impedance of a capacitor C is

ZC =

VC 1 . = IC j ωC

(7.54)

Inductors In the time domain, the voltage υL across an inductor L is related to iL by diL υL = L . (7.47) dt Phasors VL and IL are related to their time-domain counterparts by υL = Re[VL ej ωt ]

(7.48a)

iL = Re[IL ej ωt ].

(7.48b)

Because ZL and ZC are, respectively, directly and inversely proportional to ω, ZL and ZC assume inverse roles as ω approaches zero and ∞.  In the phasor domain, a capacitor behaves like an open circuit at dc and like a short circuit at very high frequencies. 

and Consequently, Re[VL ej ωt ] = L

d [Re(IL ej ωt )] = Re[j ωLIL ej ωt ], dt (7.49)

which leads to VL = j ωLIL .

(7.50)

Hence, the impedance of an inductor L is ZL =

VL = j ωL. IL

We note that the impedance of a resistor is purely real, that of an inductor is purely imaginary and positive, and that of a capacitor is purely imaginary and negative (because 1/j ωC = −j/ωC). Table 7-4 provides a summary of the υ–i properties for R, L, and C.

Example 7-4: Phasor Quantities

(7.51)

According to Eq. (7.51), ZL is positive and entirely imaginary (no real component); ZL → 0 as ω → 0 (dc); and ZL → ∞ as ω → ∞. Consequently:  In the phasor domain, an inductor behaves like a short circuit at dc and like an open circuit at very high frequencies. 

Determine the phasor-domain counterparts of the following quantities: (a) υ1 (t) = 10 cos(2 × 104 t + 53◦ ) V, (b) υ2 (t) = −6 sin(3 × 103 t − 15◦ ) V, (c) L = 0.4 mH at 1 kHz, (d) C = 2 μF at 1 MHz.

“book” — 2015/5/4 — 7:17 — page 399 — #15

7-3

PHASOR DOMAIN

399

Table 7-4: Summary of υ–i properties for R, L, and C. R

Property

L

C di dt

υ–i

υ = Ri

υ=L

V–I

V = RI

V = j ωLI

Z

R

j ωL

dc equivalent

R

Short circuit

Open circuit

Open circuit

Short circuit

|ZL|

|ZC|

R

High-frequency equivalent

|ZR| Frequency response

dυ dt I V= j ωC 1 j ωC i=C

R ω

Solution: (a) Since υ1 (t) is already in cosine format, ◦

V1 = 10ej 53 = 10 53◦ V. (b) To determine the phasor V2 corresponding to υ2 (t), we should either convert the expression for υ2 (t) to standard cosine format or apply the transformation for a sine function given in Table 7-3. We choose the first option,

1/ωC

ωL ω

ω

(c) ZL = j ωL = j 2π × 103 × 0.4 × 10−3 = j 2.5 . (d) ZC =

−j −j = −j 0.08 . = 6 ωC 2π × 10 × 2 × 10−6

υ2 (t) = −6 sin(3 × 103 t − 15◦ ) = −6 cos(3 × 103 t − 15◦ − 90◦ ) ◦

= −6 cos(3 × 10 t − 105 ) V. 3

Concept Question 7-5: Why is the phasor domain useful

for analyzing ac circuits? (See

)

To convert the amplitude from −6 to +6, we use Eq. (7.7d) of Table 7-1, namely

Concept Question 7-6: Differentiation in the time

− cos(x) = cos(x ± 180◦ ).

domain corresponds to what mathematical operation in the phasor domain? (See )

We can either add or subtract 180◦ from the argument of the cosine. Since the argument has a negative phase angle (−105◦ ), it is more convenient to add 180◦ . Hence, υ2 (t) = 6 cos(3 × 103 t − 105◦ + 180◦ ) = 6 cos(3 × 103 t + 75◦ ) V, and V2 = 6e

j 75◦

=6

75◦

V.

Concept Question 7-7: The unit for inductance is the

henry (H). What is the unit for the impedance ZL of an inductor? (See ) Concept Question 7-8: What type of circuit is equivalent

to the behavior of (a) an inductor at dc and (b) a capacitor at very high frequencies? (See )

“book” — 2015/5/4 — 7:17 — page 400 — #16

400

CHAPTER 7 AC ANALYSIS

Exercise 7-5: Determine the phasor counterparts of the following waveforms: (a) i1 (t) = 2 sin(6 × 103 t − 30◦ ) A, (b) i2(t) = −4 sin(1000t + 136◦) A

(a) I1 = 2 −120◦ A,

Answer:

(See

i

υs(t)

R

+ _

C

(b) I2 = 4 −134◦ A.

) Figure 7-6: RC circuit connected to an ac source.

Exercise 7-6: Obtain the time-domain waveforms (in

standard cosine format) corresponding to the following phasors at angular frequency ω = 3 × 104 rad/s: (a) V1 = (−3 + j 4) V

(b) V2 = (3 − j 4) V

V1 + V2 + · · · + Vn = 0,

)

I1 + I2 + · · · + In = 0, Exercise 7-7: At ω =

106

rad/s, the phasor voltage across and current through a certain element are given by V = 4 −20◦ V and I = 2 70◦ A. What type of element is it? Answer: Capacitor with C = 0.5 μF. (See

)

7-4 Phasor-Domain Analysis In the time domain, Kirchhoff’s voltage law states that the algebraic sum of all voltages υ1 to υn around a closed path containing n elements is zero, υ1 (t) + υ2 (t) + · · · + υn (t) = 0.

(7.55)

If V1 to Vn are respectively the phasor-domain counterparts of υ1 to υn , then Re[V1 ej ωt ] + Re[V2 ej ωt ] + · · · + Re[Vn ej ωt ] = 0, (7.56) or equivalently, Re[(V1 + V2 + · · · + Vn )ej ωt ] = 0.

(7.57)

Since ej ωt  = 0, it follows that Re[V1 + V2 + · · · + Vn ] = 0.

(7.58a)

Had we used a sine convention—instead of a cosine convention—we would have arrived at the result Im[V1 + V2 + · · · + Vn ] = 0.

(7.58c)

which states that KVL is equally applicable in the phasor domain. Similarly, KCL at a node leads to

Answer: (a) υ1(t) = 5 cos(3 × 104t + 126.87◦) V,

(b) υ2(t) = 5 cos(3 × 104t − 53.13◦) V. (See

The combination of Eqs. (7.58a)(a) and (b) asserts that

(7.58b)

(7.59)

where I1 to In are the phasor counterparts of i1 to in .  The fact that KCL and KVL are valid in the phasor domain is highly significant, because it implies that the analysis tools we developed earlier on the basis of these two laws also are valid in the phasor domain. These include the nodal and mesh analysis methods, the Th´evenin and Norton techniques, and several others.  Revisiting these tools and learning to apply them to ac circuits is the subject of future sections in this chapter. However, we will now introduce the basic elements of the phasor analysis process through a simple example. The phasor analysis method consists of five steps. To assist us in presenting it, we use the RC circuit shown in Fig. 7-6. The voltage source is given by υs = 12 sin(ωt − 45◦ ) V, (7.60) √ with ω = 103 rad/s, R = 3 k , and C = 1 μF. Application of KVL generates the following loop equation:  1 Ri + (time domain). (7.61) i dt = υs C Our goal is to obtain a solution for i(t). In general, i(t) consists of a transient response, obtained by solving Eq. (7.61) with υs set equal to zero (as we had done previously in Chapters 5 and 6), and a steady-state response that involves the sinusoidal

“book” — 2015/5/4 — 7:17 — page 401 — #17

7-4

PHASOR-DOMAIN ANALYSIS

401

function υs (t). Our interest at present is in only the sinusoidal response, which we can obtain by solving Eq. (7.61) in the time domain, but the method of solution is somewhat cumbersome— even for such a simple circuit—on account of the sinusoidal voltage source. Alternatively, we can obtain the desired solution by applying the phasor technique, which avoids dealing with sine and cosine functions altogether.

i Step 1 Adopt Cosine Reference (Time Domain)

υs (t) = 12 sin(ωt − 45◦ )

υs(t)

C

υs(t) = 12 sin(ωt − 45o) (V)

Step 1: Adopt cosine reference All voltages and currents with known sinusoidal functions should be expressed in the standard cosine format (Section 7-1). For our RC circuit, υs (t) is the only time-varying quantity with an explicit expression, and since υs (t) is given in terms of a sine function, we need to convert it into a cosine by applying Eq. (7.7a) of Table 7-1:

+

~+−_

R

Step 2 Transfer to Phasor Domain i υ R L C

I

R

+ _ Vs

I V ZR = R ZL = jωL ZC = 1/jωC

1 jωC o

Vs = 12e−j135 (V)

= 12 cos(ωt − 45◦ − 90◦ ) = 12 cos(ωt − 135◦ ) V. (7.62) In accordance with Table 7-3, the phasor equivalent of υs (t) is Vs = 12e

−j 135◦

V.

(7.63)

Step 3 Cast Equations in Phasor Form

(

I R+

)

1 = Vs jωC

Step 2: Transform circuit to phasor domain The current i(t) in Eq. (7.61) is related to its phasor counterpart I by i(t) = Re[Ie

j ωt

].

(7.64)

As yet, we do not have an explicit expression for either i(t) or I, but we will obtain those expressions later on in Steps 4 and 5. Step 2 in Fig. 7-7 shows the RC circuit in the phasor domain, with loop current I, impedance ZR = R representing the resistance and impedance ZC = 1/j ωC representing the capacitor. The voltage source is represented by its phasor Vs .

Step 4 Solve for Unknown Variable (Phasor Domain)

I=

Vs R+

1 jωC

i(t) = Re[Ie jωt] = 6 cos(ωt −105o) (mA)

Step 5 Transform Solution Back to Time Domain

Step 3: Cast KCL and/or KVL equations in phasor domain For the circuit in Step 2 of Fig. 7-7, its loop equation is given by ZR I + ZC I = Vs , which is equivalent to  R+

1 j ωC



Figure 7-7: Five-step procedure for analyzing ac circuits using the phasor-domain technique.

(7.65)



I = 12e−j 135 .

Step 4: Solve for unknown variable (7.66)

This equation also could have been obtained by transforming Eq. (7.61) from the time domain to the phasor domain, which  entails replacing i with I, i dt with I/j ω, and υs with Vs .

Solving Eq. (7.66) for I gives ◦

I=

12e−j 135 R+

1 j ωC



=

j 12ωCe−j 135 . 1 + j ωRC

(7.67)

“book” — 2015/5/4 — 7:17 — page 402 — #18

402

CHAPTER 7 AC ANALYSIS

Using the specified values, namely R = and ω = 103 rad/s, Eq. (7.67) becomes

√ 3 k , C = 1 μF,



I=



j 12 × 103 × 10−6 e−j 135 j 12e−j 135 = √ √ mA. 1 + j 103 × 3 × 103 × 10−6 1+j 3

υs(t)

In preparation for the next step, we should convert the expression for I into polar form (Aej θ , where A is a positive real number) because it is easier to multiply or divide two complex numbers using the polar form. To that end, we should replace j in the numerator with ej π/2 and convert the denominator into polar form: √ √ 1 + j 3 = 1 + 3 ej φ = 2ej φ , where

+

~+−_

(a)

L

Vs

(b)

Hence,

υL

Time domain

I

R

√  3 −1 = 60◦ . φ = tan 1 ◦

i

R

+ _

jωL

VL

Phasor domain



12e−j 135 · ej 90 ◦ ◦ ◦ ◦ = 6ej (−135 +90 −60 ) = 6e−j 105 mA. ◦ 2ej 60 Step 5: Transform solution back to time domain

Figure 7-8: RL circuit of Example 7-5.

I=

To return to the time domain, we apply the fundamental relation between a sinusoidal function and its phasor counterpart, namely ◦

i(t) = Re[Iej ωt ] = Re[6e−j 105 ej ωt ] = 6 cos(ωt−105◦ ) mA. This concludes our demonstration of the five-step procedure of the phasor-domain analysis technique. The procedure is equally applicable for solving any linear ac circuit.

Step 2: Transform circuit to the phasor domain. Phasor-domain circuit is shown in Fig. 7-8(b), in which R remains R, L becomes j ωL, i(t) becomes I, and υs (t) becomes Vs . Step 3: Cast KVL in phasor domain. RI + j ωLI = Vs .

Example 7-5: RL Circuit

The voltage source of the circuit shown in Fig. 7-8(a) is given by υs (t) = 15 sin(4 × 104 t − 30◦ ) V. Also, R = 3 and L = 0.1 mH. Obtain an expression for the voltage across the inductor.

Step 4: Solve for unknown variable. ◦

I=

Vs 15e−j 120 = R + j ωL 3 + j 4 × 104 × 10−4 ◦

=

Solution: Step 1: Convert υs (t) to the cosine reference.



15e−j 120 15e−j 120 ◦ = = 3e−j 173.1 A. ◦ j 53.1 3 + j4 5e

The phasor voltage across the inductor is related to I by ◦

υs (t) = 15 sin(4 × 10 t − 30 ) 4

= 15 cos(4 × 104 t − 30◦ − 90◦ ) ◦

= 15 cos(4 × 10 t − 120 ) V, 4

and its corresponding phasor Vs is given by ◦

Vs = 15e−j 120 V.



VL = j ωLI = j 4 × 104 × 10−4 × 3e−j 173.1 ◦

= j 12e−j 173.1 ◦





= 12e−j 173.1 · ej 90 = 12e−j 83.1 V, ◦

where we replaced j with ej 90 .

“book” — 2015/5/4 — 7:17 — page 403 — #19

7-5

IMPEDANCE TRANSFORMATIONS

403

Step 5: Transform solution to the time domain.

and   1 . Z3 = ZL3 + ZC3 = j ωL3 − ωC3

The corresponding time-domain voltage is obtained by multiplying VL by ej ωt and then taking the real part: ◦

υL (t) = Re[VL ej ωt ] = Re[12e−j 83.1 ej 4×10 t ] 4

From these three simple examples, we observe that an impedance Z is, in general, a complex quantity composed of a real part and an imaginary part. We usually use the symbol R to represent its real part and we call it its resistance, and we use the symbol X to represent its imaginary part and we call it its reactance. Thus,

= 12 cos(4 × 104 t − 83.1◦ ) V.

Exercise 7-8: Repeat the analysis of the circuit in Example 7-4 for υs (t) = 20 cos(2 × 103 t + 60◦ ) V, R = 6 , and L = 4 mH. Answer: υL(t) = 16 cos(2 × 103t + 96.9◦) V. (See

7-5 Impedance Transformations Voltage division, current division, and the Y– transformation are among the many analysis tools we developed in Chapter 2 in connection with circuits composed solely of sources and resistors. All of these tools are based on two fundamental laws: KCL and KVL. Having established in the preceding section that KCL and KVL also are valid in the phasor domain, it follows that these simplification and transformation techniques can be used in the phasor domain as well. The fundamental difference between the two cases is that in Chapter 2 we dealt with resistors, and with voltages and currents expressed in the time domain, whereas in the phasor domain the circuit quantities are impedances and phasors. Thus, once an ac circuit has been transformed into the phasor domain, we can apply the same techniques of Chapters 2 and 3, but we do so using complex algebra. In this and the next section, we illustrate how impedance and source transformations are executed in the phasor domain. Before we do so, however, we should expand our definition of impedance to encompass more than the impedance of a single element. The three passive elements, R, L, and C, are measured in ohms, henrys, and farads. Their corresponding impedances ZR , ZL , and ZC are all measured in ohms, and are given by ZR = R,

ZL = j ωL,

ZC =

−j . ωC

(7.68)

Consider the three series combinations shown in Fig. 7-9. Application of KVL to the circuits on the left-hand side and to their counterparts leads to Z1 = ZR1 + ZL1 = R1 + j ωL1 , Z2 = ZR2 + ZC2

j = R2 − , ωC2

Z = R + j X.

)

(7.69)

Impedances Z1 and Z2 have reactances with opposite polarities. When X is positive, as in Z1 , we call Z an inductive impedance, and when X is negative, we call it a capacitive impedance. Impedance Z2 is capacitive. Impedance Z3 is purely imaginary, and it may be inductive or capacitive depending on how the magnitude of ωL compares with that of 1/ωC. Occasionally, we may need to express Z in polar form Z = |Z|ej θ ,

(7.70)

where its magnitude |Z| and phase angle θ are related to components R and X of the rectangular form by |Z| =

 +

R 2 + X2 ,

and

θ = tan−1



 X . R

(7.71)

The inverse relationships are given by R = Re[Z] = Re[|Z|ej θ ] = |Z| cos θ

(7.72a)

X = Im[Z] = Im[|Z|ej θ ] = |Z| sin θ.

(7.72b)

and

In Chapter 2, we defined the conductance G as the reciprocal of R, namely G = 1/R. The phasor analogue of G is the admittance Y, defined as Y=

1 = G + j B, Z

(7.73)

where G = Re[Y] is called the conductance of Y and B = Im[Y] is called its susceptance. The unit for Y, G, and B is the siemen (S).

“book” — 2015/5/4 — 7:17 — page 404 — #20

404

CHAPTER 7 AC ANALYSIS

I

V

R1

jωL1

I

+ _

V

+ _

Z1 = R1 + jωL1

(a) RL I

V

R2

−j/ωC2

I

+ _

V

+ _

Z2 = R2 −

j ωC2

(b) RC I

V

jωL3 −j/ωC3

I

+ _

V

+ _

(

1 Z3 = j ωL3 − ωC 3

)

(c) LC Figure 7-9: Three different, two-element, series combinations.

7-5.1

Impedances in Series and in Parallel

Is

Voltage Division

The three in-series examples of Fig. 7-9 consisted each of only two impedances. By extension, we can assert that:  N impedances connected in series (sharing the same phasor current) can be combined into a single equivalent impedance Zeq whose value is equal to the algebraic sum of the individual impedances. 

Zeq =

N 

Vs

Z1

V1 =

(

Z1 V Z1 + Z2 s

Z2

V2 =

(

Z2 V Z1 + Z2 s

+ _

)

)

Figure 7-10: Voltage division among two impedances in series. Zi

(impedances in series).

(7.74)

i=1

The phasor voltage across any individual impedance Zi is a proportionate fraction (Zi /Zeq ) of the phasor voltage across the entire group.

This is a statement of voltage division, which for the twoimpedance circuit of Fig. 7-10, assumes the form     Z1 Z2 Vs , Vs . (7.75) V2 = V1 = Z1 + Z 2 Z1 + Z 2

“book” — 2015/5/4 — 7:17 — page 405 — #21

7-5

IMPEDANCE TRANSFORMATIONS

Is

405

Current Division I1

Vs

+ _

I1 =

(

Since Z1 = 1/Y1 and Z2 = 1/Y2 , Eq. (7.77) can be rewritten in terms of impedances as

I2



Y1

)

Y1 I Y 1 + Y2 s

Y2

I2 =

(

I1 =

)

Y2 I Y 1 + Y2 s

Z2 Z1 + Z 2



 Is ,

I2 =

Z1 Z1 + Z 2

 Is .

(7.78)

Example 7-6: Input Impedance

Figure 7-11: Current division among two admittances in

The circuit in Fig. 7-12(a) is connected to a source given by

parallel.

υs (t) = 16 cos 106 t V.

Admittance Y is the inverse of impedance Z. That is, Y = 1/Z. Hence,

Determine (a) the input impedance of the circuit, given that R1 = 2 k , R2 = 4 k , L = 3 mH, and C = 1 nF, and (b) the voltage υ2 (t) across R2 .

 N admittances connected in parallel between a pair of nodes, all sharing the same voltage, can be combined into a single, equivalent admittance Yeq , whose value is equal to the algebraic sum of the individual admittances. 

υs(t)

Yeq =

N 

Yi

(admittances in parallel)

(7.76a)

C

R1 + _

R2

L

υ2(t)

(a) Time domain

i=1

Z1

or, equivalently,

Vs

Zeq

+ _

Z1 = R1 − ZL

Zi

j ωC

ZR2 V2 ZL = jωL ZR2 = R2

−1 N  1 = . Zi

(7.76b)

(b) Phasor domain

i=1

The phasor current flowing through any individual admittance Yi is a proportionate fraction (Yi /Yeq ) of the phasor current flowing through the entire group. The current division analogue of Eq. (7.75), defining how current splits up among two admittances connected in parallel (Fig. 7-11), is  I1 =

Y1 Y1 + Y 2



 Is ,

I2 =

Y2 Y1 + Y 2

 Is .

(7.77)

Z1 Vs

+ _

Zi

Z 2 V2

Z2 = ZL || ZR2 Zi = Z1 + Z2

(c) Combining impedances Figure 7-12: Circuit for Example 7-6.

“book” — 2015/5/4 — 7:17 — page 406 — #22

406

CHAPTER 7 AC ANALYSIS

Solution: (a) The phasor-domain equivalent circuit is shown in Fig. 7-12(b), where

i

L

R1

i2 Vs = 16, j j Z1 = R1 − = (2 − j 1) k , = 2 × 103 − 6 ωC 10 × 10−9 ZL = j ωL = j × 106 × 3 × 10−3 = j 3 k ,

υs(t)

+ + _ −

~

C

R2

(a) Time domain

and ZR2 = R2 = 4 k .

I

The parallel combination of ZL and ZR2 is denoted Z2 in Fig. 7-12(c), and it is given by

Vs

+ _

ZR1 Yi

I2

I1 ZC

Za

Zb

Zb = ZC || Za

Z2 = ZL ZR2 =

ZL ZR2 j 3 × 103 × 4 × 103 j 12 × 103 = = . ZL + ZR2 (4 + j 3) × 103 4 + j3

(b) Phasor domain

A useful “trick” for converting the expression for Z2 into the form (a + j b) is to multiply the numerator and denominator by the complex conjugate of the denominator:

I ZR1 Vs

j 12 × 103 4 − j3 Z2 = × 4 + j3 4 − j3 36 + j 48 = × 103 = (1.44 + j 1.92) k . 16 + 9

Za = R2 + jωL j ZC = − ωC

+ _

Yi

(c) Combining impedances Figure 7-13: Circuit for Example 7-7.

The input impedance Zi is equal to the sum of Z1 and Z2 , Zi = Z1 + Z2 = (2 − j 1 + 1.44 + j 1.92) × 103 = (3.44 + j 0.92) k . (b) By voltage division, V2 =

Z2 Vs (1.44 + j 1.92) × 103 × 16 ◦ = = 10.8ej 38.2 V. Z1 + Z 2 (3.44 + j 0.92) × 103

Transforming V2 to its time-domain counterpart leads to υ2 (t) = Re[V2 ej ωt ] ◦

= Re[10.8ej 38.2 ej 10 t ] = 10.8 cos(106 t + 38.2◦ ) V. 6

Example 7-7: Current Division

The circuit in Fig. 7-13(a) is connected to a source υs (t) = 4 sin(107 t + 15◦ ) V.

Determine (a) the input admittance Yi , given that R1 = 10 , R2 = 30 , L = 2 μH, and C = 10 nF, and (b) the current i2 (t) flowing through R2 . Solution: (a) We start by converting υs (t) to cosine format: υs (t) = 4 sin(107 t + 15◦ ) = 4 cos(107 t + 15◦ − 90◦ ) = 4 cos(107 t − 75◦ ) V. The corresponding phasor voltage is ◦

Vs = 4e−j 75 V, and the impedances shown in Fig. 7-13(b) are given by ZR1 = R1 = 10 , ZC =

−j −j = −j 10 , = 7 ωC 10 × 10−8

“book” — 2015/5/4 — 7:17 — page 407 — #23

7-5

IMPEDANCE TRANSFORMATIONS

407

and Za = R2 + j ωL = 30 + j 107 × 2 × 10−6 = (30 + j 20) . In Fig. 7-13(c), Zb represents the parallel combination of ZC and Za , Zb = ZC Za

Concept Question 7-10: Is it possible to construct a

circuit composed solely of capacitors and inductors such that the impedance of the overall combination has a non-zero real part? Explain. (See )

Exercise 7-9: Determine the input impedance at

(−j 10)(30 + j 20) = −j 10 + 30 + j 20 20 − j 30 (20 − j 30) (3 − j 1) = = = (3 − j 11) . 3 + j1 (3 + j 1) (3 − j 1)

ω = 105 rad/s for each of the circuits in Fig. E7.9.

2 μF Zi

0.1 mH

The input impedance is Zi = ZR1 + Zb = 10 + 3 − j 11 = (13 − j 11) ,

(a)

and its reciprocal is Yi =

13 + j 11 1 1 × = Zi 13 − j 11 13 + j 11 13 + j 11 = 169 + 121

Zi

2 μF

0.1 mH



= (4.5 + j 3.8) × 10−2 = 5.89 × 10−2 e−j 40.2 S.

(b)

(b) The current I is given by I = Vs Yi = (4e

−j 75◦

)(5.89×10−2 e

Figure E7.9 −j 40.2◦

) = 0.235e

−j 34.8◦

A.

Answer: (a) Zi = j 5 , (b) Zi = −j 10 . (See

)

By current division in Fig. 7-13(b), ZC I Za + Z C −j 10 ◦ = × 0.235e−j 34.8 30 + j 20 − j 10

I2 =



=

7-5.2 Y– Transformation



2.35e−j 34.8 · e−j 90 ◦ = 7.4 × 10−2 e−j 143.2 A. ◦ j 18.4 31.6e

The corresponding current in the time domain is ◦

i2 (t) = Re[I2 ej ωt ] = Re[7.4 × 10−2 e−j 143.2 ej 10 t ] 7

= 7.4 × 10−2 cos(107 t − 143.2◦ ) A. Concept Question 7-9: The rule for adding the

capacitances of two in-series capacitors is different from that for adding the resistances of two in-series resistors, but the rule for adding the impedances of those two inseries capacitors is the same as the rule for adding two in-series resistors. Does this pose a contradiction? Explain. (See )

The Y– transformation outlined in Section 2-4 allows us to replace a Y circuit connected to three nodes with a  circuit, or vice versa, without altering the voltages at the three nodes or the currents entering them. The same principle applies to impedances, as do the relationships between impedances Z1 to Z3 of the Y circuit (Fig. 7-14) and impedances Za to Zc of the  circuit.  →Y transformation:

Z1 =

Z b Zc , Za + Z b + Z c

(7.79a)

Z2 =

Za Zc , Za + Z b + Z c

(7.79b)

“book” — 2015/5/4 — 7:17 — page 408 — #24

408

CHAPTER 7 AC ANALYSIS

1

2 Z1

c

Example 7-8: Applying Y– Transformation

Zc

1

2

Z2 Zb

Z3

Za

3

3

(a) Y circuit

(b) ∆ circuit

Solution: (a) The  circuit connected to nodes 1, 3, and 4 can be replaced with a Y circuit, as shown in Fig. 7-15(b), with impedances

Figure 7-14: Y– equivalent circuits.

Z3 =

Z a Zb . Za + Z b + Z c

Z b Zc Za + Z b + Z c −j 6 × 12 −j 72 = = = (0.8 − j 1.6) , 24 − j 12 − j 6 + 12 36 − j 18 Za Zc (24 − j 12) × 12 Z2 = = 8 , = Za + Z b + Z c 36 − j 18 and Z b Za −j 6(24 − j 12) = = −j 4 . Z3 = Za + Z b + Z c 36 − j 18 Z1 =

(7.79c)

Y→  transformation: Z1 Z2 + Z2 Z3 + Z1 Z3 Za = , Z1

(a) Simplify the circuit in Fig. 7-15(a) by applying the Y– transformation so as to determine the current I. (b) Determine the corresponding i(t), given that the oscillation frequency of the voltage source is 1 MHz.

(7.80a)

In Fig. 7-15(c), Zf represents the series combination of Z3 and Zd , Zf = Z3 + Zd = −j 4 + j 2 = −j 2 .

Zb =

Zc =

Z1 Z2 + Z2 Z3 + Z1 Z3 , Z2

Z1 Z2 + Z2 Z3 + Z1 Z3 . Z3

Similarly, (7.80b)

Zg = Z2 + Ze = (8 + j 6) . Impedances Zf and Zg are connected in parallel, and their combination is in series with Z0 and Z1 . Hence,

(7.80c)

I=

Balanced circuits:

=

If the Y circuit is balanced (all of its impedances are equal), so will be the  circuit, and vice versa. Accordingly:

Vs Z0 + Z1 + (Zf Zg ) 16ej 30



−j 2 × (8 + j 6) 2.4 + (0.8 − j 1.6) + −j 2 + 8 + j 6

.

After a few steps of complex algebra, we obtain the result Z1 = Z2 = Z3 =

Za , if Za = Zb = Zc , 3

I = 3.06 76.55◦ A.

(7.81a) (b)

Za = Zb = Zc = 3Z1 , if Z1 = Z2 = Z3 .

(7.81b)



i(t) = Re[Iej ωt ] = Re[3.06ej 76.55 ej 2π×10 t ] 6

= 3.06 cos(2π × 106 t + 76.55◦ ) A.

“book” — 2015/5/4 — 7:17 — page 409 — #25

7-5

IMPEDANCE TRANSFORMATIONS

I

Z0 = 2.4 Ω

409

1

Zb = −j6 Ω Vs

+ _

Zc = 12 Ω 24 Ω

−j12 Ω

123

3

4

Za = (24 − j12) Ω

Zd = j2 Ω Vs = 16

(a)

30o

Ze = j6 Ω 2

(V)

I

Z0 = 2.4 Ω

1 Z0 = 2.4 Ω

I

1

Z1 = (0.8 − j1.6) Ω Z1 = (0.8 − j1.6) Ω

c Vs

+ _

Z3 = −j4 Ω

Z2 = 8 Ω

3

4

Zd = j2 Ω

Ze = j6 Ω

Vs

c

+ _ Zf = Z3 + Zd = −j2 Ω

30o

Vs = 16 (b)

Vs = 16

30o

2

(V)

Zg = Z2 + Ze = (8 + j6) Ω

4

(V)

(c) Figure 7-15: Example 7-8 circuit evolution.

Exercise 7-10: Convert the Y-impedance circuit in Fig. E7.10 into a -impedance circuit.

Answer:

1

1 j7.5 Ω

j5 Ω

−j15 Ω

2 j5 Ω

−j10 Ω

2

3 Figure E7.10

3 −j15 Ω

(See

)

“book” — 2015/5/4 — 7:17 — page 410 — #26

410

CHAPTER 7 AC ANALYSIS

I +

Vs

+ -_

V12

(a)

IL

1

Zs

ZL

(a)

2

Voltage source

Actual circuit

External circuit

IL

Is

I

ZTh

1 +

Is

(b)

Zs

V12

Current source

VTh

External circuit

2

Actual circuit

7-6 Equivalent Circuits

7-6.1

(c)

 A voltage source Vs in series with a source impedance Zs is equivalent to the combination of a current source Is = Vs /Zs , in parallel with a shunt impedance Zs . The direction of Is is the same as the arrow from the (−) terminal to the (+) terminal of Vs .  Equivalence implies that both input circuits would deliver the same current I and voltage V12 to the external circuit.

7-6.2 Th´evenin Equivalent Circuit When restated for the phasor domain, Th´evenin’s theorem of Section 3-5.1 becomes:

+ Voc



VTh = Voc

Actual circuit with independent sources deactivated

Source Transformation

Section 2-3.4 provides an outline of the source-transformation principle as it applies to resistive circuits. Its phasor-domain analogue is diagrammed in Fig. 7-16 from the vantage point of the external circuit.

ZL

' Thevenin equivalent

(b)

Figure 7-16: Source-transformation equivalency.

Having examined in the preceding section how phasordomain circuits can be simplified by applying impedance transformations, we now extend our review of the rules of circuit equivalency to circuits containing voltage and current sources.

+ -_

(d)

Zeq

ZTh = Zeq

Figure 7-17: Th´evenin-equivalent method for a circuit with no dependent sources.

A linear circuit can be represented at its output terminals by an equivalent circuit consisting of a series combination of a voltage source VTh and an impedance ZTh , where VTh is the open-circuit voltage at those terminals (no load) and ZTh is the equivalent impedance between the same terminals when all independent sources in the circuit have been deactivated.  Equivalence implies that if a load ZL is connected at the output terminals of any actual circuit (as portrayed in Fig. 7-17(a)) thereby inducing a current IL to flow through it, the Th´evenin

“book” — 2015/5/4 — 7:17 — page 411 — #27

7-6

EQUIVALENT CIRCUITS

411 External-source method

Actual circuit (a)

+ Voc −

ZTh =

+ _

Vex , Iex

(7.84)

where Iex is the current generated by an external source Vex connected at the circuit’s terminals (as shown in Fig. 7-18(b)) after deactivating all independent sources in the circuit. For the sake of completeness, we should remind the reader that a Th´evenin equivalent circuit always can be transformed into a Norton equivalent circuit—or vice versa—by applying the source-transformation method of Section 7-6.1.

ZTh = Voc /Isc

Circuit with only independent sources deactivated (b)

Isc

Actual circuit

Iex Vex

ZTh = Vex /Iex

Figure 7-18: The (a) open-circuit/short-circuit method and (b) the external-source method are both suitable for determining ZTh , whether or not the circuit contains dependent sources.

Example 7-9: Thevenin ´ Circuit

The circuit shown in Fig. 7-19(a) contains a sinusoidal source given by υs (t) = 10 cos 105 t V.

equivalent circuit (Fig. 7-17(b)) would deliver the same current IL when connected to the same load impedance ZL . For the equivalence to hold, the voltage VTh and impedance ZTh of the Th´evenin circuit have to be related to the actual circuit by (Figs. 7-17(c) and (d)):

Determine the Th´evenin equivalent circuit at terminals (a, b). Solution: Step 1: The phasor counterpart of υs (t) is Vs = 10 V.

VTh = Voc

(7.82a)

ZTh = Zeq .

(7.82b)

and

Application of Eq. (7.82a) to determine VTh by calculating or measuring the open-circuit voltage Voc is always a valid approach, whether or not the actual circuit contains dependent sources. That is not so for Eq. (7.82b). The equivalentimpedance method cannot be used to determine ZTh if the circuit contains dependent sources. Alternative approaches include the following. Open-circuit / short-circuit method ZTh =

Voc , Isc

Figure 7-19(b) displays the circuit in the phasor domain, in addition to having replaced the series combination (Vs , Rs ) with the parallel combination (Is , Rs ), where Is =

Vs 10 = = 2 A. Rs 5

Step 2: Combining Rs with Z1 in parallel gives Z 1 = Rs Z1 =

5(6 + j 8) = (3.51 + j 1.08) . 5 + 6 + j8

Step 3: Converting back to a voltage source in series with Z 1 leads to the circuit in Fig. 7-19(d), with (7.83)

where Isc is the short-circuit current at the circuit’s output terminals (Fig. 7-18(a)).

Vs = Is Z 1 = 2(3.51 + j 1.08) = (7.02 + j 2.16) V.

“book” — 2015/5/4 — 7:17 — page 412 — #28

412

CHAPTER 7 AC ANALYSIS

10−5 H

υs(t)

+ _

R1 = 6 Ω

L1 = 0.08 mH

678

Rs = 5 Ω

R3 = 2 Ω C = 1 μF b

(a) υs(t) = 10 cos 105t (V)

Zs = 6.51 + j5.08

Z2 = 3 + j4 Z1 = 6 + j8

Rs = 5

Is = 2 A

Z2 = 3 + j4

Vs

+ _

+ _

RTh = 8.42 Ω

Z3 = 2 − j10

b

υTh(t)

+ _

CTh = 6.29 μF

a

7.6 cos (105t − 31.61o) V

' (f) Thevenin equivalent

b

Z2 = 3 + j4

Z3

(e) Zs = Z1 + Z2

a

(c) Z1 = Rs || Z1 Z1 = 3.51 + j1.08

Vs

Z3 = 2 − j10 b

Z1 = 3.51 + j1.08

a

a

(b) Is = Vs /Rs = 10/5 = 2 A

Is = 2 A

a

678

678

R2 = 3 Ω L2 = 4

b

a Z3 = 2 − j10

(d) Vs = IsZ1 = (7.02 + j2.16) V

b

Figure 7-19: Using source transformation to simplify the circuit of Example 7-9. (All impedances are in ohms.)

Step 4: Combining Z 1 with Z2 in series leads to the circuit in Fig. 7-19(e), where Z s

=

Z 1

+ Z2

= (3.51 + j 1.08) + (3 + j 4) = (6.51 + j 5.08) .

Step 5: Application of voltage division provides

VTh = Voc =

Vs Z3 (7.02 + j 2.16)(2 − j 10) = Z s + Z3 (6.51 + j 5.08) + (2 − j 10) = 7.6 −31.61◦ V.

“book” — 2015/5/4 — 7:17 — page 413 — #29

7-7

PHASOR DIAGRAMS

413

7-7

Hence, ◦

υTh (t) = Re[VTh ej ωt ] = Re[7.6e−j 31.61 ej 10 t ] 5



= 7.6 cos(10 t − 31.61 ) V. 5

Consider the following sinusoidal signal υs (t) and its phasor counterpart Vs :

Step 6: Suppressing the source Vs in Fig. 7-19(e) reduces the circuit at terminals (a, b) to Z s in parallel with Z3 , leading to ZTh = Z s Z3 =

(6.51 + j 5.08)(2 − j 10) = (8.42 − j 1.59) . (6.51 + j 5.08) + (2 − j 10)

Step 7: The impedance ZTh is capacitive because the sign of the imaginary component is negative. Hence, it is equivalent to ZTh = RTh −

j . ωCTh

Matching the two expressions gives RTh = 8.42 ,

CTh =

υs (t) = V0 cos(ωt + φ)

Vs = V0 φ.

(7.85)

The time-domain voltage υs (t) is characterized by three attributes: the amplitude V0 , the angular frequency ω, and the phase angle φ. In contrast, its counterpart in the phasor domain Vs is specified by only two attributes, V0 and φ. This may suggest that ω becomes irrelevant when we analyze a circuit in the phasor domain, but that certainly is not true if the circuit contains capacitors and/or inductors. Whereas ω does not appear explicitly in the expressions for phasor currents and voltages, it is integral to the definitions of the capacitor impedance ZC and inductor impedance ZL , which in turn define the I–V relationships for those two elements as

1 = 6.29 μF. 1.59ω

The time-domain Th´evenin equivalent circuit is shown in Fig. 7-19(f).

Phasor Diagrams

ZC =

VC 1 1 = = −90◦ IC j ωC ωC

(7.86a)

ZL =

VL = j ωL = ωL 90◦ . IL

(7.86b)

and

Concept Question 7-11: In the phasor domain, is the Th´evenin equivalent method valid for circuits containing dependent sources? If yes, what methods are amenable to finding ZTh of such circuits? (See )

In fact, the value of ω (relative to the values of L of C) can drastically change the behavior of a circuit:

Concept Question 7-12: If ZTh of a certain circuit is

 At dc, ZC → ∞ (open circuit) and ZL → 0 (short circuit); and conversely, as ω → ∞, ZC → 0 and ZL → ∞. 

purely imaginary, what would be your expectation about whether or not the circuit contains resistors? (See ) Exercise 7-11: Determine VTh and ZTh for the circuit in

Fig. E7.11 at terminals (a, b).

I

(10 + j30) Ω

a

Z1 10 V

+ _

Z2 5 Ω

5I b

Figure E7.11

VTh = 6 −36.9◦ V, ZTh = (2.6 + j 1.8) .

Answer:

(See

)

A phasor diagram is a useful graphical tool for examining the relationships among the various currents and voltages in a circuit. Before considering multielement circuits, however, we will start by examining the phasor diagrams for R, L and C, individually. Figure 7-20 displays the phasor diagrams for I and V for all three elements, with V chosen as a reference by selecting its phase angle to be zero. Each phasor quantity is displayed in the complex plane in terms of its magnitude and phase angle. For the resistor, VR and IR always line up along the same direction because they are always in-phase. Since VR was chosen to be purely real, so is IR . Next, we consider the capacitor. In view of Eq. (7.86a), IC =

VC = j ωCVC = ωCVC 90◦ , ZC

(7.87a)

“book” — 2015/5/4 — 7:17 — page 414 — #30

414

CHAPTER 7 AC ANALYSIS Consequently,

Resistor

 IL lags VL by 90◦ . 

Im IR VR

IR IR =

For individual elements, the relationship between I and V is straightforward; given the position of either one of them in the complex plane, we can place the other one in accordance with the phase-angle shift appropriate to that element.

R

VR

Re

VR (independent of ω) R

 For a multielement circuit, we can draw either a relative phasor diagram or an absolute phasor diagram. For the relative phasor diagram, we usually choose a specific current or voltage and designate it as our reference phasor by arbitrarily assigning it a phase angle of 0◦ . 

Capacitor Im

The goal then is to use the phasor diagram to examine the relationships between and among the various currents and voltages in the circuit—which includes their magnitudes and relative phase angles—rather than to establish their absolute phase angles. In principle, it does not matter much which specific phasor voltage or current is selected as the reference, but in practice, we usually choose a phasor current or voltage that is common to lots of elements in the circuit. By way of illustration, Example 7-10 examines a series RLC circuit by displaying its phasor diagram twice, once using the current flowing through the loop as reference, and a second time with the voltage source as reference. The former results in a relative phasor diagram, whereas the latter results in an absolute phasor diagram.

IC

IC

C

VC

90o

Re VC

IC = jωC VC (directly proportional to ω)

Inductor Im IL VL

−90o IL

L

VL

Example 7-10: Relative versus Absolute Phasor Diagrams

Re

The circuit in Fig. 7-21(a) is driven by a voltage source given by

−jVL ωL (inversely proportional to ω) IL =

υs (t) = 20 cos(500t + 30◦ ) V. Generate: (a) a relative phasor diagram by selecting the phasor current I as a reference, and (b) an absolute phasor diagram.

Figure 7-20: Phasor diagrams for R, L, and C. which positions the vector IC ahead of VC by

90◦ .

Hence:

 IC leads VC by 90◦ . 

(a) Relative Phasor Diagram Selecting I as the reference phasor means that we assign it an unknown magnitude I0 and a phase angle of 0◦ :

For the inductor, IL =

−j VL VL VL = = −90◦ . j ωL ωL ωL

Solution: Figure 7-21(b) displays the phasor-domain circuit with its RLC elements represented by their respective impedances.

(7.87b)

I = I0 0◦ .

“book” — 2015/5/4 — 7:17 — page 415 — #31

7-7

PHASOR DIAGRAMS

R=8Ω υs(t)

415

i

Relative Phasor Diagram Im

+ + _ −

~

C = 0.25 mF

2I0

VR = 8I0

Re

−2I0

I

VL + VC −3I0

VR

+ _

36.87o

−I0

(a) Time domain

Vs

ω 4I0

I = I0

L = 4 mH ZR = 8 Ω

90o

VL = 2I0

VC

−4I0

ZC = −j8 Ω

−5I0

VL

Vs = 10I0

−6I0

ZL = j2 Ω

−7I0 −8I0 VC = 8I0

(b) Phasor domain

−36.87o

−90o

(c) Relative phasor diagram where all phase angles are relative to that of I.

Im 16

VR = 16

66.87o

14

VL + VC

12 Vs = 20

10

30o

8 Absolute Phasor Diagram

6 4 VL = 4

156.87o

2

−12 −10 −8 −6 −4 −2 −2

I

66.87o 30o 2

4

6

8 10 12 14 16 18 20 −23.13o

Re

−4 −6

VC = 16 −23.13o

(d) Absolute phasor diagram Figure 7-21: Circuit and phasor diagrams for Example 7-10. The true phase angle of I is 66.87◦ , so if the relative phasor diagram in (c) were to be rotated counterclockwise by that angle and the scale adjusted to incorporate the fact I0 = 2, the diagram would coincide with the absolute phasor diagram in (d).

Because the true phase angle of I actually may not be zero, the vectors we will draw in the complex plane of the relative phasor diagram all will be shifted in orientation by exactly the same

amount (namely by the true phase angle of I) so even though they may not have the correct orientations, they all will bear the correct relative orientations to one another.

“book” — 2015/5/4 — 7:17 — page 416 — #32

416

CHAPTER 7 AC ANALYSIS

We deduce from the functional form of υs (t) that ω = 500 rad/s. In terms of I, the voltages across R, C, and L are VR = RI = 8I0 0◦ , VC =

I −j I0 = = −j 8I0 = 8I0 −90◦ , j ωC 500 × 2.5 × 10−4

Concept Question 7-14: What is the difference between a relative phasor diagram and an absolute phasor diagram? (See ) Exercise 7-12: Establish the relative phasor diagram for the circuit in Fig. E7.12 with V as the reference phasor.

and

V

VL = j ωLI = j 500 × 4 × 10−3 I0 = j 2I0 = 2I0 90◦ ,  I0 = 1 0 A

and the sum of all three gives

I1

I2

Y1 = 0.4 S

Y2 = j0.6 S

Vs = VR + VC + VL = 8I0 − j 8I0 + j 2I0  = (8 − j 6)I0 = 82 + 62 I0 ej φ = 10I0 φ,

Figure E7.12 Answer:

with φ = − tan−1

Im

6 = −36.87◦ . 8

Figure 7-21(c) displays the relative phasor diagram of the RLC circuit with I as a reference; the magnitudes of VR , VC , VL , and Vs are all measured in units of I0 , and their orientations are relative to that of I.

56.3o

(b) Absolute Phasor Diagram The phasor counterpart of υs (t) is

(See

and the application of KVL around the loop leads to Vs j R + j ωL − ωC ◦



Re I1 = 0.4V

Vs = 20 30◦ V,

I=

I0 = I1 + I2

I2 = j0.6V



20ej 30 20ej 30 20ej 30 j 66.87◦ = = A, = ◦ = 2e −j 36.87 8 + j2 − j8 8 − j6 10e

7-8

V

)

Phase-Shift Circuits

In certain communication and signal-processing applications, we often need to shift the phase of an ac signal by adding (or subtracting) a phase angle of a specified value, φ. Thus, if the input voltage in Fig. 7-22 is υin (t) = V1 cos ωt,

(7.88)

which states that the true phase angle of I is 66.87◦ . Given I, we easily can calculate VR , VC , and VL . The phasor diagram shown in Fig. 7-21(d) is identical to that in Fig. 7-21(c), except that all vectors have been rotated in a counterclockwise direction by 66.87◦ . Concept Question 7-13: For a capacitor, what is the phase angle of its phasor current, relative to that of its phasor voltage? (See )

+ υin(t) = V1 cos ωt

_

+ Phase-shift υout(t) = V2 cos(ωt + ϕ) circuit

_

Figure 7-22: The phase-shift circuit changes the phase of the input signal by φ.

“book” — 2015/5/4 — 7:17 — page 417 — #33

7-8

PHASE-SHIFT CIRCUITS

417

the function of the phase-shift circuit is to provide an output voltage given by υout (t) = V2 cos(ωt + φ).

(7.89)

For ω = 106 rad/s, R = 2 , C = 0.2 μF, and Vin = 10 V, Vout1 = 3.71 68.2◦ = (1.38 + j 3.45) V and Vout2 = 9.28 −21.8◦ = (8.62 − j 3.45) V.

The amplitude V2 of the output voltage is related to V1 (the amplitude of the input voltage) and to the configuration of the phase-shift circuit. RC circuits can be designed as phase shifters, with any specified positive or negative value of φ: 

υout leads υin υout lags υin

if 0 ≤ φ ≤ 180◦ , if − 180◦ ≤ φ ≤ 0.

To illustrate the process, let us consider the simple RC circuit shown in Fig. 7-23(a). The input signal is given by υin (t) = 10 cos 106 t

υout1 (t) = Re[Vout1 ej ωt ] = 3.716 cos(106 t + 68.2◦ ) V (7.92) and υout2 (t) = Re[Vout2 ej ωt ] = 9.285 cos(106 t − 21.8◦ ) V. (7.93)

V,

and the element values are R = 2 and C = 0.2 μF. At ω = 106 rad/s, the capacitor impedance is ZC =

The phase angle φ1 associated with Vout1 is 68.2◦ , and the angle φ2 associated with Vout2 is −21.8◦ . As shown in the complex plane of Fig. 7-23(c), the angular separation between Vout1 and Vout2 is exactly 90◦ . Also, if we were to add Vout1 and Vout2 in the complex plane, their imaginary parts would cancel out and their real parts would add up to 10 V (the amplitude of Vin ). In the time domain,

−j −j = −j 5 . = 6 ωC 10 × 0.2 × 10−6

Figure 7-23(a) provides a comparison of the waveform of the input signal υin (t) with that of υout2 (t), the voltage across the capacitor. We note that because υout2 lags υin , it always crosses the time axis later than υin by a time delay t. If we denote t0 as the time when υin (t) crosses the time axis and t2 as the time when υout2 (t) does, then ωt0 = 106 t0 =

By voltage division in the phasor domain (Fig. 7-23(b)), Vin R ωRC Vin φ1 , =√ 2 R2 C 2 j 1 + ω R− ωC   −j Vin 1 ωC = Vin φ2 , =√ j 1 + ω2 R 2 C 2 R− ωC

Vout1 =

Vout2

and

φ1 = tan−1

1 ωRC

φ2 = φ1 − 90◦ = tan−1

1 ωRC



 π  = −0.38 radians. 180◦ Now that all quantities are in the same units, we can determine the time delay from φ2 = −21.8◦ ×

(7.90b)

t2 = t2 − t0 = −φ2 × 10−6 = −(−0.38) × 10−6 = 0.38 μs.

(7.91a) 

By the same argument, υout1 leads υin by 68.2◦ , and it crosses the time axis sooner than does υin (t) by t1 = 68.2◦ ×

− 90◦ .

π , 2

with



and

ωt2 + φ2 = 106t2 + φ2 =

(7.90a)

and the phase angles φ1 and φ2 are given by 

π 2

(7.91b)

π × 10−6 = 1.19 μs. 180◦

From the foregoing analysis, we conclude that for the simple RC circuit, we can use υout1 as our output if we want to add

“book” — 2015/5/4 — 7:17 — page 418 — #34

418

CHAPTER 7 AC ANALYSIS

υa = υin(t)

i

a

t0

Input

+ R

υout1 ∆t

_b

~

υin(t) + −

υout2

t2

Lags input by ∆t2 = 0.38 μs

+ C

υout2

υb = υout2(t)

υout1 (not displayed)

_

(a) Time-domain waveforms

Leads input by ∆t1 = 1.19 μs

Im 5V 3.45 V

I

+ R=2Ω Vin

+ _

ϕ1

Vout1

_

+ ZC = −j5 Ω

Vout2

_

(b) Phasor-domain circuit

Vout1 8.62 V Vin ϕ2

−3.45 V −5 V

Re

Vout2

(c) Phasors Vin , Vout1 , and Vout2 in the complex plane

Figure 7-23: RC phase-shift circuit: the phase of υout1 (across R) leads the phase of υin (t), whereas the phase of υout2 (across C) lags the phase of υin (t).

a positive phase angle to the input υin , and we can use υout2 as our output if we want to add a negative phase angle to υin . Moreover, by adjusting the values of R and C (at a specific value of ω), we can change φ1 to any value between 0 and 90◦ , and similarly, we can change φ2 to any value between 0 and −90◦ (but not independently); as was noted earlier in connection with

Fig. 7-23(c), the absolute values of φ1 and φ2 always add up to 90◦ . Another consideration that we should be aware of is that the magnitudes of υout1 and υout2 are linked to the magnitudes of φ1 and φ2 through the choices we make for R, C, and ω. For example, as φ1 approaches 90◦ , υout1 approaches zero, so we can indeed phase-shift the input signal by an angle close to 90◦ ,

“book” — 2015/5/4 — 7:17 — page 419 — #35

7-8

PHASE-SHIFT CIRCUITS

C υs

υ1

419

C

υ2

C

Simultaneous solution of Eqs. (7.94) and (7.95), followed by several steps of algebra, leads to the expressions

υ3

+

+ _

R

R

Stage 1

R

Stage 2

υout

_

Stage 3

Figure 7-24: Three-stage, cascaded, RC phase-shifter

V1 x[(x 2 − 1) − j 3x] = 3 , Vs (x − 5x) + j (1 − 6x 2 )

(7.97)

V2 x 2 (x − j 1) = 3 , Vs (x − 5x) + j (1 − 6x 2 )

(7.98)

V3 x3 = 3 , Vs (x − 5x) + j (1 − 6x 2 )

(7.99)

x = ωRC.

(7.100)

and

(Example 7-11).

where but the magnitude of the output signal will be too small to be useful. To overcome this limitation or to introduce phase-shift angles greater than 90◦ , we can use circuits with more than two elements, such as the cascaded circuit of Example 7-11.  To generate a phase lead at the output, the cascading arrangement should be as that shown in Fig. 7-24, but to generate a phase lag, the locations of R and C should be interchanged. 

The magnitude and phase of V3 (both relative to those of Vs ) are    V3  x3  =  V  [(x 3 − 5x)2 + (1 − 6x 2 )2 ]1/2 , s and φ3 = − tan

The circuit in Fig. 7-24 uses a 3-stage cascaded phase-shifter to produce an output signal υout (t) whose phase is 120◦ ahead of the input signal υs (t). If ω = 103 (rad/s) and C = 1 μF, determine R and the ratio of the amplitude of υout to that of υs . Solution: Application of nodal analysis at nodes V1 and V2 in the phasor domain gives (7.94)

V2 − V1 V2 V2 + = 0, + R R + ZC ZC

(7.95)

where ZC = 1/j ωC. Moreover, through voltage division, V3 is related to V2 by  V2 .

 .

(7.101b)

which leads to x = 1.1815.

(7.102)

Given that ω = 103 rad/s and C = 1 μF, it follows that R=

x 1.1815 = 3 = 1.1815 k ≈ 1.2 k . ωC 10 × 10−6    V3    = 0.194. V  s

and

V3 =

1 − 6x 2 x 3 − 5x

With x = 1.1815, Eq. (7.101a) gives

V1 V1 − V2 V1 − Vs + =0 + ZC R ZC

R R + ZC



To satisfy the stated requirement, we set φ3 = 120◦ and solve for x:   1 − 6x 2 tan 120◦ = −1.732 = − , x 3 − 5x

Example 7-11: Cascaded Phase-Shifter



−1

(7.101a)

(7.96)

Note that: • The use of multiple stages allowed us to shift the phase by more than 90◦ . • However, the magnitude of the output voltage is about 20% of that of the input.

“book” — 2015/5/4 — 7:17 — page 420 — #36

420

CHAPTER 7 AC ANALYSIS

R3

Concept Question 7-15: Describe the function of a

phase-shift circuit in terms of time delay or time advance of the waveform. (See )

Exercise 7-13: Repeat Example 7-11, but use only two stages of RC phase shifters.

2Ω R1

R4





R6

2Ω iL

C

0.25 mF

L

υs2

1 mH

2Ω + _

)

R3

Exercise 7-14: Design a two-stage RC phase shifter that provides a phase shift of negative 120◦ at ω = 104 rad/s. Assume C = 1 μF.

I1

Answer: R ≈ 220 . (See

V1

)

7-9 Phasor-Domain Analysis Techniques The analysis techniques introduced in Chapter 3 in connection with resistive circuits are all equally applicable for analyzing ac circuits in the phasor domain. The only fundamental difference is that after transferring the circuit from the time domain to the phasor domain, the operations conducted in the phasor domain involve the use of complex algebra, as opposed to just real numbers. Otherwise, the circuit laws and methods of solution are identical. At this stage, instead of repeating the details of these various techniques, a more effective approach is to illustrate their implementation procedures through concrete examples. Examples 7-12 through 7-16 are designed to do just that.

R1

I2

R2

I3



2Ω I4 V2

I8 R5

I5

I6

R4





I7

C

−j4 Ω

j1 Ω

V3

I9 R6

IL −j6 V

2Ω + V s2 _

Figure 7-25: Circuit for Example 7-12 in (a) the time domain and (b) the phasor domain.

Nodal-analysis method Our first step is to transform the given circuit to the phasor domain. Accordingly, ZC =

Apply the nodal-analysis method to determine iL (t) in the circuit of Fig. 7-25(a). The sources are given by:

−j 1 = 3 = −j 4 , j ωC 10 × 0.25 × 10−3

ZL = j ωL = j 103 × 10−3 = j 1 , υs1 = 12 cos 103 t

υs1 (t) = 12 cos 103 t V,

Solution: We first demonstrate how to solve this problem using the standard nodal-analysis method (Section 3-2), and then we solve it again by applying the by-inspection method (Section 3-4).

V s1



Example 7-12: Nodal Analysis

υs2 (t) = 6 sin 103 t V.

12 V _ +

Answer: R ≈ 2.2 k ; |Vout /Vs| = 0.63. (See

R5

+

Concept Question 7-16: When is it necessary to use multiple stages to achieve the desired phase shift? (See )

υs1 2 Ω _

R2

Vs1 = 12 V,

and υs2 = 6 sin 103 t

Vs2 = −j 6 V,

where for Vs2 we used the property given in Table 7-2, namely that the phasor counterpart of sin ωt is −j . Using these values, we generate the phasor-domain circuit given in Fig. 7-25(b) in

“book” — 2015/5/4 — 7:17 — page 421 — #37

7-9

PHASOR-DOMAIN ANALYSIS TECHNIQUES

which we selected one of the extraordinary nodes as a ground node and assigned phasor voltages V1 to V3 to the other three. Our plan is to write the voltage node equations at nodes 1 to 3, solve them simultaneously to find V1 to V3 , and then use the value of V2 to obtain IL . The final step will involve transforming IL to the time domain to obtain iL (t). At node 1, KCL requires that I1 + I2 + I3 = 0.

421 which can be simplified to −V1 + (2.8 − j 0.4)V2 − V3 = 12

V3 − V 1 V3 + j 6 V3 − V2 + + = 0, 2 2 2

In terms of node voltages V1 to V3 , V1 − V3 V1 − V3 = , R3 2 V1 − V2 + Vs1 V1 − V2 + 12 I2 = , = R2 2

(7.107)

and at node 3,

(7.103)

I1 =

(node 2),

or −V1 − V2 + 3V3 = −j 6

(node 3).

(7.108)

and I3 =

Equations (7.106) to (7.108) now are ready to be cast in matrix form:

V1 V1 = . R1 + Z C 3 − j4

Inserting the expressions for I1 to I3 in Eq. (7.103) and then rearranging the terms leads to 

 1 1 1 1 1 + + V1 − V2 − V3 = −6. 2 2 3 − j4 2 2

(7.104)

The coefficient of V1 can be simplified as follows: 1 1 1 1 + + =1+ 2 2 3 − j4 3 − j4 3 − j4 + 1 = 3 − j4 4 − j4 3 + j4 = × 3 − j4 3 + j4 (12 + 16) + j (16 − 12) = = 1.12 + j 0.16. 9 + 16 (7.105) Inserting Eq. (7.105) in Eq. (7.104) and multiplying all terms by 2 leads to the following simplified algebraic equation for node 1: (2.24 + j 0.32)V1 − V2 − V3 = −12

(node 1). (7.106)

Similarly, at node 2, V2 V 2 − V3 V2 − V1 − 12 + + = 0, 2 2 + j1 2

⎤ ⎤⎡ ⎤ ⎡ −12 (2.24 + j 0.32) −1 −1 V1 ⎣ −1 (2.8 − j 0.4) −1⎦ ⎣V2 ⎦ = ⎣ 12 ⎦ . −j 6 V3 −1 −1 3 (7.109) Matrix inversion, either manually or by MATLAB or MathScript software, provides the solution: ⎡

V1 = −(4.72 + j 0.88) V,

(7.110a)

V2 = (2.46 − j 0.89) V,

(7.110b)

V3 = −(0.76 + j 2.59) V.

(7.110c)

and

Hence,

IL =

V2 2.46 − j 0.89 ◦ = = 0.81−j 0.85 = 1.17e−j 46.5 A, 2 + j1 2 + j1

and its corresponding time-domain counterpart is iL (t) = Re[IL ej 1000t ] ◦

= Re[1.17e−j 46.4 ej 1000t ] = 1.17 cos(1000t − 46.5◦ ) A. (7.111)

“book” — 2015/5/4 — 7:17 — page 422 — #38

422

CHAPTER 7 AC ANALYSIS

Y3 Y2 = V1

1 2

S

1 2

S Y5

V2

1 2

Y1

YC

1 3

j

S

Y4

6A

1 2

S

S −j3 A

IL 1 4

S = Y = (0.12+ j0.16) S

YL

V3

Y6

1 2

S

−j1 S = Y = (0.4 + j0.2) S

Figure 7-26: Equivalent of the circuit in Fig. 7-25, after source transformation of voltage sources into current sources and replacement of passive elements with their equivalent admittances.

By-inspection method Implementation of the nodal-analysis by-inspection method requires that the circuit contain no dependent sources and that all independent sources in the circuit be current sources. The first condition is valid for the circuit in Fig. 7-25(b), but the second one is not. However, both voltage sources in Fig. 7-25(b) have in-series resistors associated with them, so we easily can transform them into current sources. The resultant circuit is shown in Fig. 7-26, in which not only have the voltage sources been replaced with equivalent current sources, but all impedances have also been replaced with their equivalent admittances (Y = 1/Z). For the 3-node case, the phasor-domain equivalent of Eq. (3.25) is given by

where Y is the sum of Y1 and YC . The rule for adding two inseries admittances is the same as that for adding two in-parallel impedances: Y = Y1 YC =

+ j 41

= (0.12 + j 0.16) S.

Y11 = (1.12 + j 0.16) S. Similarly, Y22 = Y + 0.5 + 0.5 = (Y4 YL ) + 1 =

(7.112)

0.5 × (−j 1) + 1 = (1.4 − j 0.2) S, 0.5 − j 1

Y33 = 0.5 + 0.5 + 0.5 = 1.5 S.

where Ykk = Yk =

sum of all admittances connected to node k Y k = negative of admittance(s) connecting nodes k and , with k  =

Vk = unknown phasor voltage at node k Itk = total of phasor current sources entering node k (a negative sign applies to a current source leaving the node).

For the circuit in Fig. 7-26, Y11 = Y + Y2 + Y3 = (Y + 0.5 + 0.5) S,

× j 41

Hence,



⎤⎡ ⎤ ⎡ ⎤ Y11 Y12 Y13 V1 It1 ⎣Y21 Y22 Y23 ⎦ ⎣V2 ⎦ = ⎣It2 ⎦ , Y31 Y32 Y33 V3 It3

1 3 1 3

(7.113)

Also, Y12 = Y21 = Y13 = Y31 = Y23 = Y32 = −0.5 S, It1 = −6 A, It2 = 6 A, and It3 = −j 3 A. Entering the values of all of these quantities in Eq. (7.112) gives ⎤ ⎤⎡ ⎤ ⎡ −6 (1.12 + j 0.16) −0.5 −0.5 V1 ⎣ −0.5 (1.4 − j 0.2) −0.5⎦ ⎣V2 ⎦ = ⎣ 6 ⎦ . −j 3 −0.5 −0.5 1.5 V3 (7.114) Multiplication of both sides of Eq. (7.114) by a factor of 2 would produce exactly the matrix equation given by Eq. (7.109), as expected. Consequently, the final expression for iL (t) is identical to that given by Eq. (7.111). ⎡

“book” — 2015/5/4 — 7:17 — page 423 — #39

TECHNOLOGY BRIEF 19: CRYSTAL OSCILLATORS

Technology Brief 19 Crystal Oscillators Circuits that produce well-defined ac oscillations are fundamental to many applications: frequency generators for radio transmitters, filters for radio receivers, and processor clocks, among many. An oscillator is a circuit that takes a dc input and produces an ac output at a desired frequency. Temperature stability, long lifetime, and little frequency drift over time are important considerations when designing oscillators. A circuit consisting of an inductor and a capacitor √ will resonate at a specific natural frequency ω0 = 1/ LC . In such a circuit, energy is stored in the capacitor’s electric field and the inductor’s magnetic field. Once energy is introduced into the circuit (for example, by applying an initial voltage to the capacitor), it will begin to flow back and forth (oscillate) between the two components; this constant conversion gives rise to oscillations in voltage and current at the resonant frequency. In an ideal circuit with no dissipation (no resistor), the oscillations will continue at this one frequency forever. Making oscillating circuits from individual inductor and capacitor components, however, is relatively impractical and yields devices with poor reproducibility, high temperature drift (i.e., the resonant frequency changes with the temperature surrounding the circuit), and poor overall lifetime. Since the early part of the 20th century, resonators have been made in a completely different way, namely by using tiny, mechanically resonating pieces of quartz glass.

423 used in loudspeakers. Piezoelectricity can also be applied to make a quartz crystal resonate. If a voltage of the proper polarity is applied across one of the principal axes of the crystal, it will shrink along the direction of that axis. Upon removing the voltage, the crystal will try to restore its shape to its original unstressed state by stretching itself, but its stored compression energy is sufficient to allow it to stretch beyond the unstressed state, thereby generating a voltage whose polarity is opposite of that of the original voltage that was used to compress it. This induced voltage will cause it to shrink, and the process will continue back and forth until the energy initially introduced by the external voltage source is totally dissipated. The behavior of the crystal is akin to an underdamped RLC circuit. In addition to crystals, some metals and ceramics are also used for making oscillators. Because the resonant frequency can be chosen by specifying the type of material and its shape, such oscillators are easy to manufacture in large quantities, and their oscillation frequencies can be designed with a high degree of precision. Moreover, quartz crystals have good temperature performance, which means that they can be used in many applications without the need for temperature compensation, including in clocks, radios, and cellphones.

X1

Quartz Crystals and Piezoelectricity (a)

In 1880, the Curie brothers demonstrated that certain crystals—such as quartz, topaz, and tourmaline— become electrically polarized when subjected to mechanical stress. That is, such a crystal exhibits a voltage across it if compressed, and a voltage of opposite polarity if stretched. The converse property, namely that if a voltage is applied across a crystal it will change its shape (compress or stretch), was predicted a year later by Gabriel Lippman (who received the 1908 Nobel Prize in physics for producing the first color photographic plate). Collectively, these bidirectional properties of crystals are known as piezoelectricity. Piezoelectric crystals are used in microphones to convert mechanical vibrations of the crystal surface, caused by acoustic waves, into electrical signals, and the converse is

+

_

υcrystal RS

CS

LS RS = 50 Ω LS = 80 mH

(b)

CS = 1.3 fF

CO

+

υout

_

CO = 4.5 pF

Figure TF19-1: (a) Quartz crystal circuit symbol and (b) equivalent circuit. Values given are for a 5 MHz crystal.

“book” — 2015/5/4 — 7:17 — page 424 — #40

424

TECHNOLOGY BRIEF 19: CRYSTAL OSCILLATORS

Positive feedback

X1 + VCC

+

+

_

_ VCC

gain

υout

_

Negative feedback

Figure TF19-2: Schematic block diagram of an oscillator circuit. An oscillator is wired into the positive feedback path, while a negative feedback path is used to control gain.

Crystal Equivalent Circuit and Oscillator Design The electrical behavior of a quartz crystal can be modeled as a series RLC circuit (LS , CS , RS ) in parallel with a shunt capacitor (CO ). The RLC circuit models the fundamental oscillator behavior with dissipation. The shunt capacitor is mostly due to the capacitance between the two plates that actuate the quartz crystal. Figure TF19-1 shows the circuit symbol, the equivalent circuit with sample values

for a commercial 12 MHz crystal along with expressions and values for the resonant frequencies and Q. The crystal is, of course, not sufficient to produce a continuous oscillating waveform; we need to excite the circuit and keep it running. A common way to do this is to insert the crystal in the positive feedback path of an amplifier (Fig. TF19-2). The amplifier, of course, is − supplied with dc power (V+ CC and VCC ). Note that no input signal is applied to the circuit. Initially, the output generates no oscillations; however, any noise at vout that is at the resonant frequency of X1 will be fed back to the input and amplified. This positive feedback will quickly ramp up the output so that it is oscillating at the resonant frequency of the crystal. A negative feedback loop is also commonly used to control the overall gain and prevent the circuit from clipping the signal against the op amp’s − supply voltages V+ CC and VCC . In order to oscillate continuously, a circuit must meet the following two Barkhausen criteria: (1) The gain of the circuit must be greater than 1. (This makes sense, for otherwise the signal will neither get amplified nor establish a resonating condition.) (2) The phase shift from the input to the output, then across the feedback loop to the input must be 0. (This also makes sense, since if there is nonzero phase shift, the signals will destructively interfere and the oscillator will not be able to start up.)

Advances in Resonators and Clocks

Figure TF19-3: Schematic (left) and photo (right) of a tiny atomic physics package used in a chip-scale atomic clock. (Courtesy of Clark Nguyen, U.C. Berkeley, and John Kitching, National Institutes of Standards and Technology.

As good as quartz resonators are, even the best among them will drift in frequency by 0.01 ppm per year as a result of aging of the crystal. If the oscillator is being used to keep time (as in your digital watch), this dictates how many seconds (or fractions thereof) the clock will lose per year. Put differently, this drift puts a hard limit on how long a clock can run without calibration. The same phenomenon limits how well independent clocks can stay synchronized with each other. Atomic clocks provide an extra level of precision by basing their oscillations on atomic transitions; these clocks are accurate to about 10−9 seconds per day. Recently, a chip-scale version of an atomic clock (Fig. TF19-3) was demonstrated by the National Institute for Standards and Technology (NIST); it consumes 75 mW and was the size of a grain of rice (10 mm3 ). Other recent efforts for making oscillators for communication have focused on replacing the quartz crystal with a type of micromechanical resonator.

“book” — 2015/5/4 — 7:17 — page 425 — #41

7-9

PHASOR-DOMAIN ANALYSIS TECHNIQUES

425

Exercise 7-15: Write down the node-voltage matrix

or

V2 I1 V1 − 4 V1 + + + = 0, 2 j4 4 2

equation for the circuit in Fig. E7.15.

4

60o

and we also incorporate the auxiliary equation relating the two nodes, namely

(A)

V1

V2 − V1 = 29.

V2 −j4 S

I1 =

V1 − 4 . 2

(7.117)

Using Eqs. (7.116) and (7.117) in Eq. (7.115) and then solving for V1 leads to

Figure E7.15

V1 = −(4 + j 1) V,

Answer:



   ◦ (2 + j 2) −(2 + j 2) V1 2 − 4ej 60 = . ◦ −(2 + j 2) (2 − j 2) V2 4ej 60

(See

(7.116)

From the circuit, the current I1 in Eq. (7.115) is given by

(2 + j2) S

2A

(7.115)

which in turn gives IL =

)

V1 (4 + j 1) =− = (−0.25 + j 1) = 1.03 104◦ A. j4 j4

With ω = 2 ×103 rad/s, the inductor current in the time domain is given by ◦

Example 7-13: Circuit with a Supernode

iL (t) = Re[IL ej ωt ] = Re[1.03ej 104 ej 2×10 t ]

The circuit in Fig. 7-27, which is already in the phasor domain, contains two independent voltage sources, both oscillating at an angular frequency ω = 2×103 rad/s, and both characterized by a phase angle of 0◦ . Determine iL (t). Solution: Because nodes V1 and V2 are connected by a voltage source, their combination constitutes a supernode. When we apply KCL to a supernode, we simply sum all the currents leaving both of its nodes as if the two nodes are one, I1 + I2 + I3 + I4 = 0,

V1

Vs2 = 29 V V2 _

I4

+

2Ω I1

I2

+ _ V s1 = 4 V L

j4

I3 4Ω

I1 2

IL

Figure 7-27: Phasor-domain circuit containing a supernode and a dependent source (Example 7-13).

3

= 1.03 cos(2 × 103 t + 104◦ ) A.

Example 7-14: Mesh Analysis

Apply the mesh-analysis method to determine iL(t) in the circuit of Fig. 7-28, given that ω = 1000 rad/s. Solution: The circuit shown in Fig. 7-28 has mesh currents I1 to I3. Since the circuit has no dependent sources and no independent current sources, it is suitable for application of the mesh-analysis by-inspection method. For a three-loop circuit, the phasor-domain parallel of Eq. (3.28) assumes the form: ⎡ ⎤⎡ ⎤ ⎡ ⎤ Z11 Z12 Z13 I1 Vt1 ⎣Z21 Z22 Z23 ⎦ ⎣I2 ⎦ = ⎣Vt2 ⎦ , (7.118) Z31 Z32 Z33 I3 Vt3 where Zkk = sum of all impedances in loop k Zk = Z k = negative of impedance(s) shared by loop k and , with k =

Ik = unknown phasor current of loop k Vtk = total of phasor voltage sources contained in loop k, with the polarity defined as positive if Ik flows from (−) to (+) through the source.

“book” — 2015/5/4 — 7:17 — page 426 — #42

426

CHAPTER 7 AC ANALYSIS



Answer:





I3

12 V _

    (5 + j 6) −(3 + j 6) I1 12 = . −(3 + j 6) (7 + j 6) −j 6 I2



+

2Ω 3Ω

IL

I1 −j4 Ω

(See

2Ω I2 −j6 V

j1 Ω

+ _

Example 7-15: Source Superposition

The circuit in Fig. 7-29(a) contains two independent sources. Apply the source-superposition method to demonstrate that IL is given by the same expression obtained in Example 7-14, namely Eq. (7.120).

Figure 7-28: Circuit for Example 7-14.

In view of these definitions, the matrix equation for the circuit in Fig. 7-28 is given by ⎤ ⎡ ⎤⎡ ⎤ ⎡ 12 (7 − j 3) − (2 + j 1) −2 I1 ⎣−(2 + j 1) (6 + j 1) −2⎦ ⎣I2 ⎦ = ⎣ j 6 ⎦ . (7.119) −12 I3 −2 −2 6 Matrix inversion leads to I1 = (0.43 + j 0.86) A, I2 = (−0.38 + j 1.71) A, and I3 = (−1.98 + j 0.86) A.

Solution: With the source-superposition method, we activate one independent source at a time. Source 1 Alone: In part (b) of Fig. 7-29, only the 12 V source is active, and the other source has been replaced with a short circuit. The loop currents are designated I1 through I3 , and the corresponding current through the inductor is IL . Application of the mesh-current by-inspection method gives the matrix equation ⎡ ⎤⎡ ⎤ ⎡ ⎤ (7 − j 3) −(2 + j 1) −2 I1 12 ⎣−(2 + j 1) (6 + j 1) −2⎦ ⎣I ⎦ = ⎣ 0 ⎦ , (7.122) 2 −2 −2 6 I3 −12 whose inversion leads to

The current IL through the inductor is given by

I1 = (0.79 + j 0.52) A,

IL = I1 − I2 = (0.43 + j 0.86) − (−0.38 + j 1.71) ◦

= 0.81 − j 0.85 = 1.17e−j 46.5 A,

and −j 46.5◦

I3 = (−1.86 + j 0.33) A.

ej 1000t ]

= 1.17 cos(1000t − 46.5◦ ) A. (7.121) Exercise 7-16: Write down the mesh-current matrix



4Ω 3Ω

I1

j6 Ω Figure E7.16

Hence, IL = I1 − I2 = (0.79 + j 0.52) − (−0.36 + j 0.48) = (1.15 + j 0.04) A.

equation for the circuit in Fig. E7.16.

+ 12 V _

I2 = (−0.36 + j 0.48) A,

(7.120)

and its time-domain counterpart is iL (t) = Re[IL ej ωt ] = Re[1.17e

)

I2

+ _

j6 V

(7.123)

Source 2 Alone: Deactivation of the 12 V source and reactivation of the −6j V source produces the circuit shown in part (c) of Fig. 7-29. Now the loop currents are I1 , I2 , and I3 , and their matrix equation is ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ 0 (7 − j 3) −(2 + j 1) −2 I1 ⎣−(2 + j 1) (6 + j 1) −2⎦ ⎣I ⎦ = ⎣j 6⎦ . (7.124) 2 0 I3 −2 −2 6

“book” — 2015/5/4 — 7:17 — page 427 — #43

7-9

PHASOR-DOMAIN ANALYSIS TECHNIQUES

427

2Ω 12 V _





+

2Ω 3Ω



IL

−j4 Ω

j1 Ω

(a)

+ _

−j6 V

Both sources 2Ω

2Ω 12 V _



I3

+

2Ω 3Ω I1 −j4 Ω

IL

I3











I2

I1 −j4 Ω

j1 Ω (c)

(b) −j6 V source replaced with short circuit

2Ω 2Ω

IL

I2 −j6 V

j1 Ω

+ _

12 V source replaced with short circuit

Figure 7-29: Demonstration of the source-superposition technique (Example 7-15).

The solution of Eq. (7.124) is

Example 7-16: Thevenin ´ Approach

I1 = (−0.36 + j 0.34) A,

For the circuit of Fig. 7-30, (a) obtain its Th´evenin equivalent at terminals (a, b), as if the inductor were an external load, and (b) then use the Th´evenin circuit to determine IL .

I2 = (−0.02 + j 1.23) A, I3 = (−0.13 + j 0.53) A, and IL = I1 − I2 = −0.36 + j 0.34 − (−0.02 + j 1.23) = (−0.34 − j 0.89) A. Total Superposition Solution: Given IL due to source 1 alone and IL due to source 2 alone, the total current due to both sources simultaneously is IL = IL + IL = (1.15 + j 0.04) + (−0.34 − j 0.89) = (0.81 − j 0.85) A,

(7.125)

which is identical to the expression given by Eq. (7.120).

Solution: (a) We will apply the open-circuit/short-circuit method to determine the values of VTh and ZTh of the Th´evenin equivalent circuit. Open-Circuit Voltage: With the inductor replaced with an open circuit in Fig. 7-30(b), the matrix equation for loop currents I1 and I2 is      12 + j 6 (9 − j 4) −4 I1 = , (7.126) −12 I2 −4 6 and its inversion gives I1 = (0.02 + j 0.96) A

and

I2 = (−1.98 + j 0.64) A.

“book” — 2015/5/4 — 7:17 — page 428 — #44

428

CHAPTER 7 AC ANALYSIS



2Ω 2Ω



12 V _

I2

+

12 V _





+

2Ω 2Ω

IL

3Ω a

+ _

j1 Ω

−j4 Ω







a −j6 V

I1

b (a) Original circuit

+ _

Voc

−j4 Ω

−j6 V

b

(b) Inductor replaced with open circuit

2Ω 12 V _



I5

ZTh 2Ω

IL

+





3Ω I3

I4

a Isc

−j4 Ω

a

+ _

VTh

L + _

−j6 V

j1 Ω

b

b ' (d) Thevenin circuit connected to inductor

(c) Inductor replaced with short circuit

Figure 7-30: After determining the open-circuit voltage in part (b) and the short-circuit current in part (c), the Th´evenin equivalent circuit is connected to the inductor to determine IL .

With I1 and I2 known, application of KVL around the loop containing the −j 6 V source leads to

I3 = (0.44 + j 0.95) A, I4 = (−0.53 + j 1.60) A,

VTh = Voc = 2(I1 − I2 ) + 2I1 − j 6 = 4I1 − 2I2 − j 6 = (4.06 − j 3.44) V. (7.127) Short-Circuit Current: In part (c) of Fig. 7-30, the inductor has been replaced with a short circuit. The matrix equation for loop currents I3 to I5 is given by ⎤ ⎤⎡ ⎤ ⎡ 12 (7 − j 4) −2 −2 I3 ⎣ −2 6 −2⎦ ⎣I4 ⎦ = ⎣ j 6 ⎦ . −12 I5 −2 −2 6

Solution of Eq. (7.128) gives



and I5 = (−2.03 + j 0.85) A, from which we have Isc = I3 − I4 = (0.44 + j 0.95) − (−0.53 + j 1.60) = (0.97 − j 0.65) A.

(7.129)

Given Voc and Isc , it follows that (7.128)

ZTh =

Voc 4.06 − j 3.44 = = (4.53 − j 0.51) . (7.130) Isc 0.97 − j 0.65

“book” — 2015/5/4 — 7:17 — page 429 — #45

7-10 AC OP-AMP CIRCUITS

429

(b) Having established VTh and ZTh , we now connect the Th´evenin equivalent circuit to the inductor at terminals (a, b), as shown in Fig. 7-30(d). The current IL is simply IL =

4.06 − j 3.44 VTh = = (0.80 − j 0.85) A. ZTh + j 1 4.53 − j 0.51 + j 1 (7.131)

υp ip

7-10 ac Op-Amp Circuits in Question 1: Are op amps used in ac circuits?

υn

+ + Ro (υp − υn) Ri A(υp − υn) − + -_ −

+

io

υo

Answer 1: Yes. Question 2: Is the ideal op-amp model applicable to ac circuits?

To explain what we mean by the answer to the second question, let us start with a quick review of the op-amp models introduced earlier in Chapter 4. The operation of the op amp can be represented by the equivalent circuit shown in Fig. 7-31(a). The model parameters include large input resistance Ri on the order of megaohms, small output resistance Ro on the order of 50 , and an open-loop gain A. At dc, A is very large, on the order of 105 or greater. These attributes allowed us to adopt the ideal op-amp model in which we set Ri ≈ ∞, Ro ≈ 0, and A ≈ ∞. By invoking these approximations, we obtained the two constraints: υp = υn and ip = in = 0. The use of these constraints served to significantly simplify the analysis of op-amp circuits containing dc sources. An important underlying assumption is that A is very large. Whereas this assumption is certainly valid for dc, it is not necessarily so at ac.

(a) Op-amp equivalent circuit

+

υp in = 0 υn

(Ri =

_

)

8

Answer 2: The ideal op-amp model is based on the assumption that the open-loop gain A is very large (> 104 ), which is true at dc and low frequencies, but not necessarily so at high frequencies. The range of frequencies over which A is large depends on the specific op-amp design. As we shall see later on in this section, when the standard LM741 op amp is used in an inverting amplifier circuit, the ideal op-amp model is applicable for ac circuits so long as the frequency is less than about 1 kHz. For operations at higher frequencies, other models should be used instead, so the selection of a particular op-amp model for a particular application (such as amplification and processing of video signals) becomes an important consideration.

(Ro = 0)

+

υo

υp = υn

(b) Ideal op-amp model Figure 7-31: Op-amp (a) equivalent circuit (for both dc and ac) and (b) ideal model (for dc, and ac at low frequencies).

Figure 7-32 displays a typical plot of the open-loop gain A as a function of the oscillator frequency f for the LM741 op amp. At dc, the gain (denoted A0 ) is indeed very large (105 ), but A decreases rapidly with increasing frequency. The gain spectrum of an op amp is characterized by three important parameters: (a) the dc gain A0 : the value of A at f = 0 Hz. (b) the corner √ frequency fc : the frequency at which A = A0 / 2 = 0.707A0 . (c) the unity gain frequency fu : the frequency at which A = 1. For the op-amp gain displayed in Fig. 7-32, A0 = 105 , fc = 10 Hz, and fu = 1 MHz. The ideal op-amp model assumes

“book” — 2015/5/4 — 7:17 — page 430 — #46

430

CHAPTER 7 AC ANALYSIS

A

Rf = 10 kΩ dc gain 0.707 × 105

105

Rs

103

υn

2 kΩ υs

LM741

104

~ + −

υp

_ υo

+

RL 10 kΩ

102 10

(a) Inverting amplifier circuit

Corner frequency fc

1

Unity gain frequency fu

0.1 10

100

1k

Rf = 10 kΩ

f (Hz) 10 k 100 k 1 M 10 M

Rs = 2 kΩ

Figure 7-32: Open-loop gain A versus frequency for the LM741 op amp.

υs

that A is very large, which is a valid assumption at dc and at frequencies as high as 10 kHz, but it is certainly not valid at much higher frequencies. What are the implications of a nonuniform spectrum for A (i.e., A not a constant as a function of f )? We answer the question through Example 7-17.

~ + −

υn

Ro

Ri 1 MΩ A(υp − υn) + _ υp

υo RL 10 kΩ

(b) Equivalent circuit model Figure 7-33: Inverting amplifier.

Equivalent circuit model The node equations at nodes υn and υo in Fig. 7-33(b) are given by:

Example 7-17: Audio and Video Amplifier

The objective of this example is to establish whether or not the inverting amplifier circuit shown in Fig. 7-33(a) is suitable for amplifying (a) audio signals with spectra extending to 1 kHz and video signals with spectra extending to 1 MHz. The op-amp gain spectrum is given in Fig. 7-32, and the input and output resistances are Ri = 1 M , and Ro = 50 .

υn υn − υo υn − υs + + = 0, Rs Ri Rf υo − A(υp − υn ) υo υo − υn + + = 0. Rf Ro RL

(7.133) (7.134)

After setting υp = 0 (because the positive input terminal is connected to the ground terminal) and solving the two equations simultaneously to obtain an expression for the circuit gain, we have

Solution: Since A is not uniformly high at all the frequencies under consideration, we should compute the circuit gain Rf υo = G= G = υo /υs using the op-amp equivalent circuit model, and then υs Rs compare it with the value obtained using the ideal model. We   Rs Ri (Ro − ARf ) will perform the comparison at multiple frequencies between · . (RL Ro + Rf RL + Rf Ro )(Ri Rf + Rs Rf + Rs Ri ) − Rs Ri (Ro − ARf ) dc and 1 MHz. (7.135) Ideal op-amp model From Eq. (4.24), Gideal =

υo Rf 10 k = −5. =− =− υs Rs 2k

(7.132)

Using the values Ri = 106 , Rs = 2 × 103 , Rf = 104 , Ro = 50 , RL = 10 k , and the value of A from Fig. 7-32, we obtain the results summarized in Table 7-5. (a) Audio Signal: Based on the gain data listed in Table 7-5, an audio signal consisting of frequencies extending between

“book” — 2015/5/4 — 7:17 — page 431 — #47

7-11

OP-AMP PHASE SHIFTER

431

Table 7-5: Inverting amplifier gain G as a function of

Zf

oscillation frequency f . Gideal = −5

f (Hz)

A

G

0 (dc) 100 1k 10 k 100 k 1M

105 104 103 102 10 1

−4.997 −4.970 −4.714 −3.111 −0.707 −0.081

The error is defined as



% error =

Gideal − G Gideal

C2 Error 0.06% 0.6% 5.7% 37.8% 85.9% 98.4%

υin

× 100.

Figure 7-34: Inverting amplifier as a phase-shift circuit.

Zs R1

C1

R2

_ +

~+−

+

υout

_



1 j ωR1 C1 + 1 = , (7.137a) j ωC1 j ωC1 1 R2 /j ωC2 R2 Zf = R2 = = . j ωC2 R2 + 1/j ωC2 1 + j ωR2 C2 (7.137b) Zs = R1 +

dc and 1 kHz would experience relatively minimal distortion because they would all be amplified by a factor of about −5, within a maximum variation of 5.7% (at 1 kHz). (b) Video Signal: Because the video signal extends to 1 MHz and the op-amp circuit does not provide good amplification at frequencies above 10 kHz, the output signal will be highly distorted. Hence, to amplify video signals with minimal distortion, it is necessary to use an op amp with a corner frequency (Fig. 7-32) as high as 1 MHz or higher.

The circuit gain is Zf −j ωR2 C1 Vout =− = , Vin Zs (1 + j ωR1 C1 )(1 + j ωR2 C2 ) (7.138) which can be expressed as G=

7-11 Op-Amp Phase Shifter

G = |G|ej φ ,

In Section 7-8, we examined how an RC circuit can be used as a phase shifter with an output voltage having the same angular frequency ω of the input voltage, but whose phase angle is increased or decreased (shifted) by a desired amount. That is, if the input is υin (t) = V1 cos ωt,

(7.136a)

the phase-shifted output is υout (t) = V2 cos(ωt + φ).

(7.136b)

As was shown earlier in Section 7-8, an RC circuit can indeed realize the desired phase shift, but at a cost in amplitude. The amplitude of the output voltage, V2 , is smaller than V1 , and the degree of reduction depends on φ and the number of RC stages used in the phase shifter. An op-amp circuit can serve as a phase shifter, without necessarily sacrificing a reduction in amplitude. Consider the circuit in Fig. 7-34(a). It is an inverting amplifier with complex source and feedback impedances:

(7.139)

with ωR2 C1 , [(1 + ω2 R12 C12 )(1 + ω2 R22 C22 )]1/2

(7.140a)

φ = 270◦ − tan−1 (ωR1 C1 ) − tan−1 (ωR2 C2 ),

(7.140b)

|G| =

where 270◦ is the phase angle corresponding to (−j ) in the numerator of Eq. (7.138). In the time domain, υout (t) = |G|V1 cos(ωt + φ).

(7.141)

Through judicious choice of the values of R1 , R2 , C1 , and C2 , it should be possible to design a phase shifter that provides the desired value of φ, with |G| ≥ 1. The process is illustrated by Example 7-18. Example 7-18: Op-Amp Phase Shifter

Select values for R1 , R2 , C1 , and C2 in the circuit of Fig. 7-34 so that φ = 120◦ and |G| = 2 at ω = 500 rad/s.

“book” — 2015/5/4 — 7:17 — page 432 — #48

432

CHAPTER 7 AC ANALYSIS

υs(t) ac input

Transformer

Rectifier

Filter

Vout dc output

Voltage regulator

υs(t) Vout

Figure 7-35: Block diagram of a basic dc power supply.

Solution: With 4 selectable parameters against only 2 specified parameters, the desired outcome can be realized through many different combinations of (R1 , R2 , C1 , C2 ). Hence, we arbitrarily choose R1 = 2 k ,

C1 = 3 μF,

which leads to ωR1 C1 = 500 × 2 × 103 × 3 × 10−6 = 3. Using this value and φ = 120◦ in Eq. (7.140b) leads to

7-12 Application Note: Power-Supply Circuits Systems composed of one or more electronic circuits usually contain power-supply circuits that convert the ac power available from the wall outlet into dc power, thereby providing the internal dc voltages required for proper operation of the electronic circuits. Most dc power supplies consist of the four subsystems diagrammed in Fig. 7-35. The input is an ac voltage υs (t) of amplitude Vs and angular frequency ω, and the final output is a dc voltage Vout . Our plan in this section is to describe the operation of each of the intermediate stages, and then connect them all together.

120◦ = 270◦ − tan−1 (3) − tan−1 (ωR2 C2 ),

7-12.1

which simplifies to tan−1 (ωR2 C2 ) = 270◦ − 120◦ − tan−1 3 = 150◦ − 71.57◦ = 78.43◦ . Hence, ωR2 C2 = tan 78.43◦ = 4.89. With ωR1 C1 = 3 and ωR2 C2 = 4.89, and |G| = 2, solution of Eq. (7.140a) leads to R2 = 21 k ,

Ideal Transformers

A transformer consists of two inductors called windings, that are in close proximity to each other but not connected electrically. The two windings are called the primary and the secondary, as shown in Fig. 7-36. Even though the two windings are isolated electrically—meaning that no current flows between them—when an ac voltage is applied to the primary, it creates a magnetic flux that permeates both windings through a common core, inducing an ac voltage in the secondary.  The transformer gets its name from the fact that it is used to transform currents, voltages, and impedances between its primary and secondary circuits. 

and C2 =

4.89 4.89 = = 0.47 μF. ωR2 500 × 21 × 103

The key parameter that determines the relationships between the primary and the secondary is the turns ratio n = N2 /N1 ,

“book” — 2015/5/4 — 7:17 — page 433 — #49

7-12 APPLICATION NOTE: POWER-SUPPLY CIRCUITS

i1

N1 : N2 i2

i1

+

+

+

υ1

υ2

υ1

_

_

Dots on same ends

_

N1 : N2

_ υ2

+

i2 Dots on opposite ends

Figure 7-36: Schematic symbol for an ideal transformer. Note the reversal of the voltage polarity and current direction when the dot location at the secondary is moved from the top end of the coil to the bottom end. For both configurations: υ2 p2 N2 i2 N1 1 υ 2 i2 = = n, = = , = =1 υ1 N1 i1 N2 n p1 υ 1 i1

where N1 is the number of turns in the primary coil and N2 is the number of turns in the secondary. An additionally important attribute is the direction of the primary winding, relative to that of the secondary, around the common magnetic core. The relative directions determine the voltage polarity and current direction at the secondary, relative to those at the primary. To distinguish between the two cases, a dot usually is placed at one or the other end of each winding, as shown in Fig. 7-36. For the ideal transformer, voltage υ2 at the secondary side is related to voltage υ1 at the primary side by υ2 N2 = = n, υ1 N1

(7.142)

where the polarities of υ1 and υ2 are defined such that their (+) terminals are at the ends with the dots. In an ideal transformer, no power is lost in the core, so all of the power supplied by a source to its primary coil is transferred to the load connected at its secondary side. Thus, p1 = p2 , and since p1 = i1 υ1 and p2 = i2 υ2 , it follows that i2 N1 = , i1 N2

(7.143)

433 direction of magnetic flux coupling between the two coils. More details are available in Chapter 11.  If N2 /N1 > 1, the transformer is called a step-up transformer because it transforms υ1 to a higher voltage, and if N2 /N1 < 1, it is called a step-down transformer.  Most office and household electronic gadgets (such as telephones, clocks, radios, and answering machines) require dc voltages that are on the order of volts (or at most a few tens of volts), which is much smaller than the voltage level available at the wall outlet. The transformer in such gadgets is invariably a step-down transformer. As discussed in great detail in Chapter 11, the inputoutput relationships for a real transformer are more elaborate than those given by Eqs. (7.142) and (7.143) for the ideal transformer. Nevertheless, these simple relationships are reasonable first-order approximations and serve our current discussion quite adequately. Concept Question 7-17: In a transformer, how are the

voltage polarities and current directions defined relative to the dots on the primary and secondary windings? (See ) Concept Question 7-18: For an ideal transformer, how

is power p2 related to power p1? (See

7-12.2

Rectifiers

A rectifier is a diode circuit that converts an ac waveform into one that is either always positive or always negative, depending on the direction(s) of the diode(s). Power supplies usually use a bridge rectifier, but to appreciate how such a bridge functions, we will first consider the simple single-diode rectifier circuit shown in Fig. 7-37. As discussed in Section 2-6.2, a diode is modeled by a practical response that allows current to flow through it in the direction shown in Fig. 7-37 if and only if the voltage across it is greater than a threshold value known as the forward-bias voltage VF . That is, for the circuit in Fig. 7-37, the output voltage across the load resistor is given by 

with i1 always defined in the direction towards the dot on the primary side and i2 defined in the direction away from the dot on the secondary side. The purpose of the dot designation is to indicate whether the windings in the primary and secondary coils curl in the same (clockwise or counterclockwise) direction or in opposite directions. The coil directions determine the

)

υout =

υin − VF 0

if υin ≥ VF , if υin ≤ VF .

(7.144)

For an ideal diode with VF = 0, the output waveform is identical to the input waveform for the half cycles during which υin is positive, and the output is zero when υin is negative. In the case

“book” — 2015/5/4 — 7:17 — page 434 — #50

434

CHAPTER 7 AC ANALYSIS

υout(t) with VF = 0

υin(t) υin

i1

+ VF _

υout υout(t) with VF = 0.7 V

+ _

RL

υout(t)

υin(t) Figure 7-37: Half-wave rectifier circuit.

of a real diode with VF ≈ 0.7 V, the peak amplitude of the output is smaller than that of the input by 0.7 V. Because the output waveform essentially replicates only the positive half cycles of the input waveform (with a negative amplitude shift equal to VF ), the circuit of Fig. 7-37 is called a half-wave rectifier. Next, we consider the bridge-rectifier circuit of Fig. 7-38. The bridge rectifier uses four diodes. During the positive half cycle of υin (t), two of the diodes conduct, and the other two are OFF. The reverse happens during the second half cycle, but the direction of the current through RL is the same during both half cycles. Consequently, the output waveform essentially is equivalent to taking the absolute value of the input waveform (if VF is so small relative to the peak value as to be neglected). Because a bridge rectifier acts on both halves of a cycle, it is often called a full-wave rectifier. Exercise 7-17: Suppose the input voltage in the circuit of Fig. 7-38 is a 10 V amplitude square wave. What would the output look like? Answer: 8.6-V dc. (See

7-12.3

)

Smoothing Filters

So far, we have examined two of the four subcircuits of the dc power supply. The transformer serves to adjust the amplitude of the ac signal to a level close to the desired dc voltage level of the final output. The bridge rectifier converts the ac signal into an all-positive waveform. Next, we need to reduce the variations of the full-wave rectified waveform to bring it to as close to a constant level as possible. We accomplish this by subjecting the full-wave rectified waveform to a smoothing (averaging) filter. This is realized by adding a capacitor C in

parallel with the load resistor. The modified circuit is shown in Fig. 7-39(a), and the associated output waveform is displayed in Fig. 7-39(b). The capacitor is a storage device that goes through partial charging-up and discharging-down cycles. During the charging-up period, the upswing time constant of the circuit is given by τup = (2RD RL )C ≈ 2RD C

if RL  RD ,

(7.145)

where RD is the diode resistance. Typically, RD is on the order of ohms and RL is on the order of kiloohms, so the approximation given by Eq. (7.145) is quite reasonable. In the absence of the capacitor in the circuit, RD usually is ignored because it is in series with a much larger resistance, RL . Adding a capacitor, however, creates an RC circuit in which R is the parallel combination of RD and RL , placing RD in a controlling position. During the discharging period, the diode turns off, and the capacitor discharges through RL alone. Consequently, the downswing time constant involves RL and C only, τdn = RL C.

(7.146)

For a specified value of the diode resistance RD , we can choose the values of RL and C so that τup is short and τdn is long— both relative to the period of the rectified waveform—thereby realizing a fast response on the upswing part and a very slow response on the downswing part. In practice, it is possible to generate an approximately constant dc voltage with a ripple component on the order of 1 to 10 percent of its average value (Fig. 7-39(b)).

“book” — 2015/5/4 — 7:17 — page 435 — #51

7-12 APPLICATION NOTE: POWER-SUPPLY CIRCUITS

435

OFF

υin(t) υin

υout(t)

_ υ out +

+ _

RL OFF

(a) Positive half cycle

OFF

υin

υout(t)

_ υ out +

+ _

RL

υin(t) OFF

(b) Negative half cycle υout(t) = |υin(t)| − 2VF

υin(t) Full-wave bridge rectifier (c) Input-output response

Figure 7-38: Full-wave bridge rectifier. Current flows in the same direction through the load resistor for both half cycles.

Example 7-19: Filter Design

Trect =

1 = 8.33 ms, 120

If the bridge rectifier circuit of Fig. 7-39(a) has a 60 Hz ac input signal, determine the values of RL and C that would result in τup = Trect /12 and τdn = 12Trect , where Trect is the period of the rectified waveform. Assume RD = 5 .

and the corresponding design specifications are

Solution: If the frequency of the original ac signal is 60 Hz, the frequency of the rectified waveform is 120 Hz. Hence, the period of the rectified waveform is

Application of Eq. (7.145) leads to

τup =

Trect = 0.69 ms, 12

and

τdn = 12Trect = 100 ms.

τup ≈ 2RD C

“book” — 2015/5/4 — 7:17 — page 436 — #52

436

CHAPTER 7 AC ANALYSIS

υin

+ _

+ C

υout

RL

(a) Bridge rectifier with filter υout

Capacitor charging up Capacitor discharging

With filter

Without filter Ripple voltage υr

t

Trect (b) Filtered output Figure 7-39: Smoothing filter reduces the variations of waveform υout (t).

or τup 0.69 × 10−3 C= = = 69 μF. 2RD 2×5 With the value of C known, application of Eq. (7.146) gives τdn 100 × 10−3 RL = = 1.45 k . = C 69 × 10−6

7-12.4 Voltage Regulator The circuit shown in Fig. 7-40 includes all of the power-supply subcircuits we have discussed thus far, plus two additional elements, namely a series resistance Rs and a zener diode. When operated in reverse breakdown, the zener diode maintains the voltage across it at a constant level Vz —so long as the current iz passing through it remains between certain limits. Since the diode is connected in parallel with RL , the output voltage becomes equal to the zener voltage Vz , and the effective time constant of the smoothing filter becomes τ = Rs C. It is worth noting that the addition of the zener diode reduces the peak-to-peak ripple voltage Vr (Fig. 7-39(b)) at the output of

the RC filter by about an order of magnitude. An approximate expression for the peak-to-peak ripple voltage with the zener diode in place is given by Vr =

[(Vs1 − 1.4) − Vz ]Trect (Rz RL ) × , Rs C Rs + (Rz RL )

(7.147)

where Vs1 is the amplitude of the ac signal at the output of the transformer (Fig. 7-40), the factor 1.4 V accounts for the voltage drop across a pair of diodes in the rectifier, Vz is the manufacturer-rated zener voltage for the specific model used in the circuit, Trect is the period of the rectified waveform, and Rz is the manufacturer specified value of the zener-diode resistance. Example 7-20: Power-Supply Design

A power supply with the circuit configuration shown in Fig. 7-40 has the following specifications: the input voltage is 60 Hz √ with an rms amplitude Vrms = 110 V where Vrms = Vs / 2 (the rms value of a sinusoidal function is

“book” — 2015/5/4 — 7:17 — page 437 — #53

7-13

MULTISIM ANALYSIS OF AC CIRCUITS

437

υs2 Vs1 =

( )

N2 V N1 s

υout Vz

(Vs1 − 1.4) t

N1 : N2

υs(t) = Vs cos ωt

+ _

t

Zener diode Vz Rs

υs1(t) = Vs1 cos ωt

iz

+ υs2(t)

C

_ Transformer

Rectifier

+ RL υout(t)

_

123 RC filter and voltage regulator

Figure 7-40: Complete power-supply circuit.

discussed in Chapter 8), N1 /N2 = 5, C = 2 mF, Rs = 50 , RL = 1 k , Vz = 24 V, and Rz = 20 . Determine υout , the ripple voltage, and the ripple fraction relative to υout . Solution: At the secondary side of the transformer,   N2 (Vs cos 377t) υs1 (t) = N1 √ 1 = × 110 2 cos 377t = 31.11 cos 377t V. 5 Hence, Vs1 = 31.11 V, which is greater than the zener voltage Vz = 24 V. Consequently, the zener diode will limit the output voltage at υout = Vz = 24 V. In Example 7-19, we established that Trect = 8.33 ms. Also, 20 × 1000 = 19.6 . Rz RL = 20 + 1000 Application of Eq. (7.147) gives [(Vs1 − 1.4) − Vz ]Trect (Rz RL ) × Vr = Rs C Rs + (Rz RL ) 19.6 [(31.11 − 1.4) − 24] (8.33 × 10−3 ) × = 50 × 2 × 10−3 50 + 19.6 = 0.13 V (peak-to-peak). Hence, 0.13/2 (Vr /2) = = 0.0027, Vz 24 which represents a relative variation of less than ±0.3 percent. ripple fraction =

7-13

Multisim Analysis of ac Circuits

Even though we usually treat the wires in a circuit as ideal short circuits, in reality a wire has a small but non-zero resistance. Also, as noted earlier in Section 5-7.1, when two wires are in close proximity to one another, they form a non-zero capacitor. A pair of parallel wires on a circuit board is modeled as a distributed transmission line with each small length segment

represented by a series resistance R and a shunt capacitance C, as depicted by the circuit model shown in Fig. 7-41. For a parallel-wire segment of length , R and C are given by R= or R=

2

π a2σ !

πf μ σ

(low-frequency approximation) √ (a f σ ≤ 500), (7.148a) 

πa



(high-frequency approximation) √ (a f σ ≥ 1250), (7.148b)

and C=

π ε

ln(d/a)

for (d/2a)2  1,

(7.148c)

where a is the wire radius, d is the separation between the wires, f is the frequency of the signal propagating along the wires, μ and σ are respectively the magnetic permeability and conductivity of the wire material, and ε is the permittivity of the material between the two wires. Note that R represents the

“book” — 2015/5/4 — 7:17 — page 438 — #54

438

CHAPTER 7 AC ANALYSIS and if the total length of the parallel wires is t = 15 cm, then their transmission-line equivalent circuit should consist of n sections with

t 15 cm n= = = 5.

3 cm We will now use Multisim to simulate such a transmission line.

Conductivity σ 2a d l

l

l

l

Example 7-21: Transmission-Line Simulation

R

R

C

R C

R C

C

Figure 7-41: Distributed impedance model of two-wire transmission line.

A pair of parallel wires made of a conducting material with conductivity σ = 1.9×105 S/m is used to carry a 1 GHz squarewave signal between two circuits on a circuit board. The wires are 15 cm in length and separated by 1 mm, and their radii are 0.1 mm. (a) Develop a transmission-line equivalent model for the wires and (b) use Multisim to evaluate the voltage response along the transmission line. Solution: (a) With = 3 cm (to satisfy Eq. (7.149)), application of Eqs. (7.148b and c) gives

resistance of both wires. There is actually a third distributed element to consider in the general case of a transmission line: the distributed inductance. This inductance is placed in series with the resistance R of each segment. It arises because current flowing through the transmission-line wires gives rise to a magnetic field around the wires and, hence, an inductance (as discussed in Section 5-3). However, modeling the behavior of a transmission line with all three components is rather complex. So, for the purposes of this section, we will ignore the inductance altogether so that we may illustrate the performance of an RC transmission line using Multisim. Keeping this in mind, the distributed model shown in Fig. 7-41 allows us to represent the wires by a series of cascaded RC circuits. For the model to faithfully represent the behavior of the real twowire configuration, each RC stage should represent a physical length that is no longer than a fraction (≈ 10 percent) of the distance that the signal travels during one period of the signal frequency. Thus, should be on the order of



up T c ≈ , 10 10f

(7.149)

where up is the signal velocity along the wires, which is on the order of the velocity of light by c = 3 × 108 m/s, and the period T is related to the frequency f by T = 1/f . For example, if the signal frequency is 1 GHz (= 109 Hz), then should be on the order of c 3 × 108

≈ = 3 cm, = 10f 10 × 109

! R= " =

πf μ σ



πa



π × 109 × 4π × 10−7 1.9 × 105



3 × 10−2 π × 10−4



= 13.76 and C= =

π ε

ln(d/a) π × (10−9 /36π ) × 3 × 10−2 ln(10)

= 3.6 × 10−13 F = 0.36 pF. (b) To use Multisim, we need to select values for R and C— from the libraries of available values—that are approximately equal to those we calculated. The selected values are less critical to the simulation than the value of their product, because it is the product RC = 13.76 × 0.36 × 10−12 ≈ 5 × 10−12 s that determines the time constant of the voltage response. Hence, we select R = 10

and

C = 0.5 pF,

and we draw the 5-stage circuit shown in Fig. 7-42. The square wave is generated by a pulse generator that alternates between 0 and 1 V. Its pulses are 500 ps long and the pulse period is

“book” — 2015/5/4 — 7:17 — page 439 — #55

7-13

MULTISIM ANALYSIS OF AC CIRCUITS

439

Figure 7-42: Transmission-line circuit in Multisim.

1000 ps (or equivalently, 1 ns, which is the period corresponding to a frequency f = 1 GHz). The Rise Time and Fall Time should be set to 1 ps. Figure 7-43 displays V(1) at node 1, which represents the pulse-generator voltage waveform, and the voltages at nodes 2, 3, 4, 5, and 6 corresponding to the outputs of the five RC stages.

During the charging-up period, it takes longer for the nodes further away from the pulse generator to reach the steady-state voltage of 1 V than it does for those closer to the generator. The same pattern applies during the discharge period. In addition to the parallel-wire configuration, the distributed transmission-line concept is equally applicable

Input voltage V(1)

V(6) V(4) V(2)

Figure 7-43: Multisim display of voltage waveforms at nodes 1, 2, 3, 4, 5, and 6.

“book” — 2015/5/4 — 7:17 — page 440 — #56

440

CHAPTER 7 AC ANALYSIS

Figure 7-44: Using the Logic Analyzer to measure time delay in Multisim.

to other transmission media, including the shielded cable commonly used for the transmission of audio, video, and digital data between different circuits. If a digital signal with logic 0 = 0 V and logic 1 = 1 V is to be transmitted along a coaxial cable or some other transmission line, it may be of interest to simulate the process using Multisim to determine how long it takes to charge the different nodes along the line up to 1 V. This is also known as propagating the logic 1 down the transmission line. The Logic Analyzer (Simulate → Instruments → Logic Analyzer) is used to visualize a large number of logic levels at once. (See the Multisim Tutorial for a detailed explanation on how to use the Logic Analyzer Instrument.) An example is shown in Fig. 7-44. The circuit uses 1 M resistors, 5 fF capacitors, and a pulse generator. The pulse length is set at 500 ps and the pulse period at 1000 ps (= 1 ns). The circuit nodes are wired to the logic analyzer. In Fig. 7-45, we can observe how long it takes each node to charge up sufficiently to register as a logic 1. Note that the logic analyzer’s cursor can be used to read out the exact time points. Example 7-22: Measuring Phase Shift

Run a Transient Analysis on the Multisim circuit in Fig. 7-44 after replacing the pulse generator with a 1 V amplitude, 10 MHz ac source. The goal is to determine the phase of node 2,

relative to the phase of node 1 (the voltage source). Select a Start Time of 2.7 μs and an End Time (TSTOP) of 3.0 μs, and set TSTEP and TMAX to 1e-10 seconds so as to generate smooth-looking curves. [We did not choose a Start Time of 0 s simply because it takes the circuit a few microseconds to reach its steady-state solution.] Solution: Figure 7-46 shows the traces of selected nodes V(1), V(2), and V(6) on Grapher View. Clicking on the Show/Hide Cursors button enables the measurement cursor, which can be used to quantify the amplitude (vertical axis) and time (horizontal axis) for each curve. To measure the phase shift between nodes V(2) and V(1), two cursors are needed. Step 1: Place cursor 1 slightly to the left of a maximum of the V(1) trace.

Step 2: Click on the trace for V(1) to select it. White triangles will appear on the V(1) trace.

Step 3: Right-click the cursor itself and select Go to next Y Max=>. On row x1, at column V(1), the value in the table should be 2.7250 μs.

“book” — 2015/5/4 — 7:17 — page 441 — #57

7-13

MULTISIM ANALYSIS OF AC CIRCUITS

Node 1

Node 2

Node 4

Node 6

Figure 7-45: Logic Analyzer readout at nodes 1, 2, 3, 4, 5, and 6.

V(1) V(2) V(6)

Figure 7-46: Multisim Grapher Plot of voltage nodes V(1), V(2), and V(6) in the circuit of Fig. 7-42.

441

“book” — 2015/5/4 — 7:17 — page 442 — #58

442

CHAPTER 7 AC ANALYSIS

Figure 7-47: Using Measurement Probes to determine phase and amplitude of signal at various points on transmission line. Step 4: Repeat the process using cursor 2 to select the nearby maximum of the V(2) trace. The entry in row x2, at column V(2), should be 2.7312 μs. The time difference between the two values is

Exercise 7-18: Determine the amplitude and phase of V(6) in the circuit of Example 7-22, relative to those of V(1). Answer: (See

)

t = 2.7312 μs − 2.7250 μs = 0.0062 μs. Given that f = 10 MHz, the period is 1 f 1 = 7 10

T =

= 10−7 = 0.1 μs. Application of Eq. (7.11) gives 

 t φ = 2π T   0.0062 = 360◦ × 0.1 = 22.3◦ . We also can determine the ratio of the amplitude of V(2) to that of V(1). The ratio of y2 in column V(2) to y1 in column V(1) gives V(2) 0.656 = ≈ 66 percent. V(1) 1

Additional method to measure amplitude and phase Let us continue working with the transmission-line circuit of the previous two examples. Place a Measurement Probe (of the type we introduced in Chapters 2 and 3) at each of the appropriate nodes in the circuit. Double-click on the Probe, and under the Parameters tab, select the appropriate parameters so that only V(p-p), Vgain(ac), and Phase are printed in the Probe output. Additionally, with the exception of Probe 1 (located right above V1), at the top of the Probe Properties window, check Use reference probe, and select Probe 1. Note that “phase” here refers to the phase difference between the voltage at the specific probe and the reference probe. So if a particular signal is leading the reference node, then the phase will appear negative, and if a particular signal is lagging the reference node, then the phase will appear positive. This is the opposite of how we are taught to think of phase, so keep this at the front of your mind when using this approach. Run the Interactive Simulation by pressing F5 (or any of the appropriate buttons or toggles, which you should know by now) and the result should resemble that shown in Fig. 7-47. We can see that the Phase at Node 2 is 22.6◦ , which of course is opposite to what we see in Fig. 7-46, where the signal at V(2) is behind V(1) by 22.3◦ . However, we must remember that the phase values are flipped in the Measurement Probe readings, so the values actually are in agreement. Additionally, we see in Fig. 7-47 that the Vgain(ac) at Node 2 is “654m” (which corresponds to 65.4 percent), which is very nearly in agreement with the value of 66 percent obtained in Example 7-22.

“book” — 2015/5/4 — 7:17 — page 443 — #59

7-13

MULTISIM ANALYSIS OF AC CIRCUITS

443

Summary Concepts • A sinusoidal waveform is characterized by three independent parameters: its amplitude A, its angular frequency ω, and its phase angle φ. • Complex algebra is used extensively in the phasor domain to analyze ac circuits. Hence, it behooves every student taking a course in circuit analysis to become proficient in using complex numbers (by hand, with a scientific calculator, and with MATLAB/Mathworks). • By transforming an ac circuit from the time domain to the phasor domain, its integro-differential equation gets transformed into a linear equation. After solving the linear equation, the solution is then transformed back to the time domain. • Voltages and currents in the time domain have phasor



• • •



counterparts in the phasor domain; resistors, capacitors, and inductors are transformed into impedances. The rules for combining impedances (when connected in series or in parallel) are the same as those for resistors in resistive circuits. The same is true for Y– transformations. All of the techniques of circuit analysis are equally applicable in the phasor domain. A phase shifter is a circuit that can modify the phase angle of a sinusoidal waveform. An ac waveform can be converted into dc by subjecting it to a four-step process that includes a transformer, bridge rectifier, smoothing filter, and voltage regulator. Multisim is very useful for analyzing an ac circuit and evaluating its response as a function of frequency.

Mathematical and Physical Models Transformer

Trigonometric identities

Table 7-1

Time domain/phasor domain correspondence

Table 7-2

Impedance

ZR = R ZC = 1/j ωC ZL = j ωL Zeq =

Impedances in series

Admittances in parallel Y– transformation

Important Terms absolute phasor diagram ac admittance alternating current amplitude

Yeq =

Section 7-4.2

N  i=1 N 

υ2 N2 = υ1 N1 i2 N1 = i1 N2  2

for (a f σ ≤ 500) π a2σ !   πf μ

R= σ πa √ for (a f σ ≥ 1250)

Wire resistance

R=

Wire capacitor

C=

Zi

Yi

i=1

π ε

ln(d/a)

for (d/2a)2  1

Provide definitions or explain the meaning of the following terms: angular frequency argument bridge rectifier capacitive impedance complex conjugate

complex number conductance core cosine-referenced current division

downswing time constant electromagnetic compatibility Euler’s identity forward-bias voltage

“book” — 2015/5/4 — 7:17 — page 444 — #60

444

CHAPTER 7 AC ANALYSIS

Important Terms (continued) frequency frequency domain technique full-wave rectifier half-wave rectifier iff ideal transformer imaginary impedance inductive impedance lag lead oscillation frequency peak-to-peak ripple voltage

peak value period (of a cycle) phase angle phase lag phase lead phase-shift circuit phase-shift oscillator phasor counterpart phasor diagram phasor domain phasor domain technique polar form primary winding radio frequency identification

PROBLEMS Section 7-1: Sinusoidal Signals *7.1

Express the sinusoidal waveform υ(t) = −4 sin(8π × 103 t − 45◦ ) V

in standard cosine form and then determine its amplitude, frequency, period, and phase angle. 7.2

Express the current waveform i(t) = −0.2 cos(6π × 109 t + 60◦ ) mA

in standard cosine form and then determine the following: (a) Its amplitude, frequency, and phase angle. (b) i(t) at t = 0.1 ns. *7.3 A 4 kHz sinusoidal voltage waveform υ(t), with a 12 V amplitude, was observed to have a value of 6 V at t = 1 ms. Determine the functional form of υ(t). 7.4 Two waveforms, υ1 (t) and υ2 (t), have identical amplitudes and oscillate at the same frequency, but υ2 (t) lags υ1 (t) by a phase angle of 60◦ . If υ1 (t) = 4 cos(2π × 103 t + 30◦ ) V, write the expression appropriate for υ2 (t) and plot both waveforms over the time span from −1 ms to +1 ms. ∗ Answer(s) available in Appendix G.

reactance real relative phasor diagram rectangular form rectifier resistance ripple secondary winding sinusoidal waveform source-transformation principle step-down transformer step-up transformer susceptance

Th´evenin’s theorem time-independent time shift time-varying function transformer true turns ratio upswing time constant voltage division voltage regulator winding zener diode zener-diode resistance zener voltage

7.5 Waveforms υ1 (t) and υ2 (t) are given by: υ1 (t) = −4 sin(6π × 104 t + 30◦ ) V, υ2 (t) = 2 cos(6π × 104 t − 30◦ ) V. Does υ2 (t) lead or lag υ1 (t), and by what phase angle? *7.6 A phase angle of 120◦ was added to a 3 MHz signal, causing its waveform to shift by t along the time axis. In what direction did it shift and by how much? 7.7 Provide an expression for a 24 V signal that exhibits adjacent minima at t = 1.04 ms and t = 2.29 ms. 7.8 A multiplier circuit has two input ports, designated υ1 and υ2 , and one output port whose voltage υout is equal to the product of υ1 and υ2 . Assume υ1 = 10 cos 2πf1 t V, υ2 = 10 cos 2πf2 t V. (a) Obtain an expression for υout in terms of the sum and difference frequencies, fs = f1 + f2 and fd = f1 − f2 . (b) Plot its waveform over the time interval [0, 2 s], given that f1 = 3 Hz and f2 = 2 Hz. *7.9 Provide an expression for a 12 V signal that exhibits a maximum at t = 2.5 ms, followed by an adjacent minimum at t = 12.5 ms.

“book” — 2015/5/4 — 7:17 — page 445 — #61

PROBLEMS

445

Section 7-2: Complex Algebra 7.10 Express the following complex numbers in polar form: (a) z1 = 3 + j 4 (b) z2 = −6 + j 8 *(c) z3 = −6 − j 4

(g) z7 = (−1 + j )1/2 7.11 Express the following complex numbers in rectangular form: (a) z1 = 2ej π/6 (b) z2 = −3e−j π/4 √ *(c) z3 = 3 e−j 3π/4 (d) z4 = −j 3 (e) z5 = −j −4 *(f) z6 = (2 + j )2 (g) z7 = (3 − j 2)3 Complex numbers z1 and z2 are given by: z1 = 6 − j 4, z2 = −2 + j 1. (a) Express z1 and z2 in polar form. (b) Determine |z1 | by applying Eq. (7.20) to the given expression. (c) Determine the product z1 z2 in polar form. (d) Determine the ratio z1 /z2 in polar form. (e) Determine z12 and compare it with |z1 |2 . (f) Determine z1 /(z1 − z2 ) in polar form. For the complex number z = 1 + j , show that z2 − |z|2 = −2(1 − j ). 7.14

If z = −8 + j 6, determine the following quantities:

(a) |z|2 *(b) z2 , in polar form (c) 1/z, in polar form (d) z−3 , in polar form (e) Re(1/z2 )

Complex numbers z1 and z2 are given by

z2 = 5 45◦ .

(f) z6 = (3 − j 2)3

7.13

7.15

z1 = 2 −60◦ ,

(d) z4 = j 2 *(e) z5 = (2 + j )2

7.12

*(f) Im(z∗ ) (g) Im[(z∗ )2 ] (h) Re[(z∗ )−1/2 ]

Determine in polar form: (a) z1 z2 (b) z1 /z2 *(c) z1 z2∗ (d) z12 √ (e) z2  ∗ (f) z2 (g) z1 (z2 − z1 )∗ (h) z2∗ /(z1 + z2 ) 7.16 Given z = 1.2 − j 2.4, determine the value of: (a) ln z *(b) ez (c) ln(z∗ ) (d) exp(z∗ + 1) 7.17 Simplify the following expressions into the form (a + j b), where a and b are real numbers: √ √ (a) j + −j √ √ (b) j −j (1 + j )2 (c) (1 − j )2 7.18 Simplify the following expressions and express the result in polar form: ◦

5e−j 30 − j4 2 + j3 (−20 45◦ )(3 − j 4) + (2 + j ) *(b) B = (2 − j ) 1 j4 + (c) C = (3 + j 2) − 2(1 − j ) 1 + j 4    (2 − j ) −(3 + j 4)  (d) D =  −(3 + j 4) (2 + j )     5 30◦ −2 45◦   (e) E =  −2 45◦ 4 60◦  (a) A =

7.19 Calculate the following complex numbers and express the results in rectangular form:

“book” — 2015/5/4 — 7:17 — page 446 — #62

446

CHAPTER 7 AC ANALYSIS

(a) 3ej π/4 − j 4 (b) (3 − j 4)/(2 45◦ ) (c) [(7 − j 5)/(3 + j 2)] − 2ej 3π/4

(a) Apply the necessary trigonometric identities from Table 7-1 to show that

7.20 Calculate the following complex numbers and express the results in polar form: (a) 2ej π/3 − 5ej π/2 *(b) (8 − j 3)(7 + j 2) (c) [(6 − j 7)/(−2 + j 9)] + 3e−j π/4

(b) Transform the expression given by Eq. (1) to the phasor domain, simplify it into a single term, and then transform it back to the time domain to show that the result is identical to the expression given by Eq. (2).

υ(t) = 9.67 cos(ωt + 48.1◦ ) V.

Sections 7-3 to 7-5: Phasor Domain and Impedance Transformations 7.21 Transform the following sinusoidal currents into their phasor counterparts: (a) i1 (t) = 10 sin(8t + 75◦ ) A (b) i2 (t) = −17 cos(9t − 25◦ ) A (c) i3 (t) = [8 cos(6t − 45◦ ) − 5 sin(6t)] A 7.22 Determine the phasor counterparts of the following sinusoidal functions: (a) υ1 (t) = 4 cos(377t − 30◦ ) V (b) b υ2 (t) = −2 sin(8π × 104 t + 18◦ ) V (c) υ3 (t) = 3 sin(1000t + 53◦ ) − 4 cos(1000t − 17◦ ) V 7.23 Determine the instantaneous time functions corresponding to the following phasors: ◦ (a) I1 = 6ej 60 A at f = 60 Hz ◦ (b) I2 = −2e−j 30 A at f = 1 kHz *(c) I3 = j 3 A at f = 1 MHz (d) I4 = −(3 + j 4) A at f = 10 kHz (e) I5 = −4 −120◦ A at f = 3 MHz 7.24 Show that the instantaneous time function corresponding ◦ ◦ to the phasor V = 4ej 60 + 6e−j 60 V is given by υ(t) = 5.29 cos(ωt − 19.1◦ ) V. 7.25 Determine the impedances of the following elements: (a) R = 1 k at 1 MHz (b) L = 30 μH at 1 MHz *(c) C = 50 μF at 1 kHz

7.27 Use phasors to simplify each of the following expressions into a single term [Hint: See Problem 7.26]: (a) υ1 (t) = 12 cos(6t + 30◦ ) − 6 cos(6t − 45◦ ) V *(b) υ2 (t) = −3 sin(1000t − 15◦ ) − 6 sin(1000t + 15◦ ) + 12 cos(1000t − 60◦ ) V (c) υ3 (t) = 2 cos(377t + 60◦ ) − 2 cos(377t − 60◦ ) V (d) υ4 (t) = 10 cos 800t + 10 sin 800t V 7.28 Simply the following expressions using phasors: (a) i1 (t) = 20 cos(ωt − 30◦ ) + 16 cos(ωt + 15◦ ) A (b) i2 (t) = 14 sin(ωt + 45◦ ) − 17 cos(ωt + 60◦ ) A (c) i3 (t) = 2 cos(5t) − 7 sin(5t) A *7.29 The current source in the circuit of Fig. P7.18 is given by is (t) = 12 cos(2π × 104 t − 60◦ ) mA. Apply the phasor-domain analysis technique to determine iC (t), given that R = 20 and C = 1 μF.

iC is(t)

R

C

Figure P7.29: Circuit for Problems 7.29 and 7.30.

7.30 Repeat Problem 7.29, after replacing the capacitor with a 0.5 mH inductor and then calculating the current through it. 7.31 Find is (t) in the circuit of Fig. P7.31, given that υs (t) = 15 cos(5 × 104 t − 30◦ ) V, R = 1 k , L = 120 mH, and C = 5 nF.

is(t)

υs(t)

+ _

R

C

7.26 The function υ(t) is the sum of two sinusoids, υ(t) = 4 cos(ωt + 30◦ ) + 6 cos(ωt + 60◦ ) V.

(2)

(1) Figure P7.31: Circuit for Problem 7.31.

L

“book” — 2015/5/4 — 7:17 — page 447 — #63

PROBLEMS

447

*7.32 Find voltage υab (t) in the circuit of Fig. P7.32, given that is (t) = 35 sin(300t − 15◦ ) mA, R = 80 , L = 15 mH, and C = 200 μF.

50 Ω Z1

10 mH

10 μF

a

+ is(t)

R

(a) C

υab(t)

0.3 H

1 kΩ

L

_

Z2

b

0.15 H

(b)

Figure P7.32: Circuit for Problem 7.32.

50 Ω 20 nF

7.33 Find ia (t) in the circuit of Fig. P7.33, given that υs (t) = 40 sin(200t − 20◦ ) V.

Z3

200 μF

0.2 μF

25 nF

30 Ω

10 Ω (c)

υs(t)

ia(t)

+ _

10 Ω 0.5 μF

30 Ω 50 mH

10 Ω Z4

Figure P7.33: Circuit for Problem 7.33.

0.1 mH

10 Ω

Z

j3 Ω



50 μF

20 μF 20 Ω

(e)

j6 Ω −j2 Ω

30 Ω 10 mH

Z5

Find the input impedance Z of the circuit in Fig. P7.35.

−j3 Ω

10 Ω

(d)

7.34 Determine the equivalent impedance: (a) Z1 at 1000 Hz (Fig. P7.34(a)) (b) b Z2 at 500 Hz (Fig. P7.34(b)) (c) Z3 at ω = 106 rad/s (Fig. P7.34(c)) (d) Z4 at ω = 105 rad/s (Fig. P7.34(d)) (e) Z5 at ω = 2000 rad/s (Fig. P7.34(e)) 7.35

1 μF

Figure P7.34: Circuits for Problem 7.34.

5Ω a



3 mH

Figure P7.35: Circuit for Problem 7.35.

Z *7.36 Find the input impedance Z of the circuit in Fig. P7.36 at ω = 400 rad/s.

2 mF



b Figure P7.36: Circuit for Problem 7.36.

9 mH

“book” — 2015/5/4 — 7:17 — page 448 — #64

448 7.37

CHAPTER 7 AC ANALYSIS Find Z of the circuit in Fig. P7.37, given that ◦

υs (t) = 10 cos(377t + 15 ) V

*7.40 The circuit in Fig. P7.40 is in the phasor domain. Determine the following: (a) The equivalent input impedance Z at terminals (a, b).

and ◦

(b) The phasor current I, given that Vs = 25 45◦ V.

is (t) = 3 sin(377t + 30 ) A.

Is I + _

Vs



a

Z + _

Vs

7.41

IR

−j5 Ω j12 Ω

10 Ω

Use the phasor domain circuit in Fig. P7.41.

(a) Determine the value of Zx that would make the input impedance Z purely real. (b) Specify what type of element would be needed to realize that condition, and what its magnitude should be if ω = 6250 rad/s.

Figure P7.38: Circuit for Problem 7.38.

a

7.39 The voltage source in the circuit of Fig. P7.39 is given by υs (t) = 12 cos 104 t V. (a) Transform the circuit to the phasor domain and then determine the equivalent impedance Z at terminals (a, b). [Hint: Application of –Y transformation should prove helpful.] (b) Determine the phasor I, corresponding to i(t). (c) Determine i(t).

+ υs(t) _

a Z b



0.3 mH

0.3 mH



Figure P7.39: Circuit for Problem 7.39.

Zx j1 Ω

j1 Ω



1Ω 2Ω

Z

j1 Ω

j1 Ω b

50 μF i(t)

j5 Ω

Figure P7.40: Circuit for Problem 7.40.

13 Ω

+ _

j5 Ω b

Find IR in the circuit of Fig. P7.38, given that Vs = 25 V.

Vs

j5 Ω

Z

Figure P7.37: Circuit for Problem 7.37.

*7.38

j5 Ω −j5 Ω





Figure P7.41: Circuit for Problem 7.41.

3Ω 7.42 In response to an input signal voltage υs (t) = 24 cos 2000π t, the input current in the circuit of Fig. P7.42 was measured as i(t) = 6 cos(2000π t − 60◦ ) mA. Determine the equivalent input impedance Z of the circuit.

“book” — 2015/5/4 — 7:17 — page 449 — #65

PROBLEMS

449 Sections 7-6: Equivalent Circuits

i(t)

a

+ υs(t) _

Z

Passive circuit

b

Figure P7.42: Configuration for Problem 7.42.

7.43 At ω = 400 rad/s, the input impedance of the circuit in Fig. P7.43 is Z = (74 + j 72) . What is the value of L?

7.46 Your objective is to obtain a Th´evenin equivalent for the circuit shown in Fig. P7.46, given that is (t) = 3 cos 4 × 104 t A. To that end: (a) Transform the circuit to the phasor domain. (b) Apply the source-transformation technique to obtain the Th´evenin equivalent circuit at terminals (a, b). (c) Transform the phasor-domain Th´evenin circuit back to the time domain.

1 mH

Ω

a

10 Ω

a

is(t)

25 Ω 20 Ω

Z

b Figure P7.46: Circuit for Problem 7.46.

7.47 The input circuit shown in Fig. P7.47 contains two sources, given by

Figure P7.43: Circuit for Problem 7.43.

is (t) = 2 cos 103 t A,

*7.44 In the circuit of Fig. P7.44, what should the value of L be at ω = 104 rad/s so that i(t) is in-phase with υs (t)?

i(t)

a

50 Ω 25 Ω

Z

0.5 μF

L

50 μF

b

+ υs(t) _

350 37

35 Ω

4 μF

L

b

υs (t) = 8 sin 103 t V. This input circuit is to be connected to a load circuit that provides optimum performance when the impedance Z of the input circuit is purely real. The circuit includes a “matching” element whose type and magnitude should be chosen to realize that condition. What should those attributes be?

Matching element a ?

Figure P7.44: Circuit for Problem 7.44.

6Ω 7.45 At what angular frequency ω is the current i(t) in the circuit of Fig. P7.45 in-phase with the source voltage υs (t)?

is(t)

2Ω 0.5 mF

4 mH

+ _ υs(t)

Z

Load circuit

b i(t) υs(t)

+ _

a

108.33 μF

50 Ω Figure P7.47: Circuit for Problem 7.47.

0.2 H

Z b

0.2 mF

0.2 mF *7.48 Determine the Th´evenin equivalent of the circuit in Fig. P7.48 at terminals (a, b), given that υs (t) = 12 cos 2500t V,

Figure P7.45: Circuit for Problem 7.45.

is (t) = 0.5 cos(2500t − 30◦ ) A.

“book” — 2015/5/4 — 7:17 — page 450 — #66

450

CHAPTER 7 AC ANALYSIS

3 kΩ

5Ω 4 mH + _ υs(t)

4 mH

80 μF

a is(t)

Figure P7.48: Circuit for Problem 7.48.

0.1 Ω

a

0.2 Ω 0.2 Ω

+ _ 0.2I

7.52 For the circuit in Fig. P7.52: (a) Apply current division to express IC and IR in terms of Is . (b) Using Is as reference, generate a relative phasor diagram showing IC , IR , and Is and demonstrate that the vector sum IR + IC = Is is satisfied. (c) Analyze the circuit to determine Is and then generate the absolute phasor diagram with IC , IR , and Is drawn according to their true phase angles.

b

+ _ 15

2000Ix +_

a

Ix ZTh

−j6 kΩ 0

V

ZL

45o

V

−j3 Ω



Figure P7.52: Circuit for Problem 7.52.

7.53 For the circuit in Fig. P7.53: (a) Apply current division to express I1 and I2 in terms of Is . (b) With Is as reference, generate a relative phasor diagram showing that the vector sum I1 +I2 = Is is indeed satisfied. (c) Analyze the circuit to determine Is and then generate the absolute phasor diagram for the three currents.

b

Figure P7.50: Circuit for Problem 7.50.

*7.51 The phasor current IL in the circuit of Fig. P7.51 was measured to be   78 36 IL = +j mA. 41 41

IR

IC

+ _ Vs = 2



Determine ZL .

Is



7.50 As we will learn in Chapter 8, to maximize the transfer of power from an input circuit to a load ZL , it is necessary to choose ZL such that it is equal to the complex conjugate of the impedance of the input circuit. For the circuit in Fig. P7.50, such a condition translates into requiring ZL = Z∗Th . Determine ZL such that it satisfies this condition.

3 kΩ

ZL

Figure P7.51: Circuit for Problem 7.51.

Figure P7.49: Circuit for Problem 7.49.

4 kΩ

j6 kΩ

Sections 7-7 and 7-8: Phasor Diagrams and Phase Shifters

7.49 The circuit in Fig. P7.49 is in the phasor domain. Determine its Th´evenin equivalent at terminals (a, b).

j0.1 Ω

IL

+ 15 0 V _

10 Ω b

I

2 kΩ

+ _ Vs = 10

0

V

Is I1

I2





j4 Ω

Figure P7.53: Circuit for Problem 7.53.

−j3 Ω

“book” — 2015/5/4 — 7:17 — page 451 — #67

PROBLEMS

451

7.55 A two-stage RC circuit provides a phase shift lead of 120◦ . What is the ratio of the output-voltage amplitude to that of the input?

2A −j6 V +

V1

Section 7-9: Analysis Techniques 7.57 Apply nodal analysis in the phasor domain to determine ix (t) in the circuit of Fig. P7.57.

V3

IC

j0.2 Ω

*7.56 The element values of a single-stage phase-shift circuit are R = 40 and C = 5 μF.At what frequency f is φ1 = −φ2 , where φ1 and φ2 are the phase angles of the output voltages across R and C, respectively?

0.4 Ω

V2

_

7.54 Design a two-stage 1 MHz RC phase-shift circuit whose output voltage is 120◦ behind that of the input signal. All capacitors are 1 nF each. Determine the values of the resistors and the amplitude ratio of the output voltage to that of the input.

4A

−j0.5 Ω

Figure P7.59: Circuit for Problem 7.59.

7.60 Apply the by-inspection method to develop a nodevoltage matrix equation for the circuit in Fig. P7.60, and then use MATLAB or MathScript software to solve for V1 and V2.

j12 Ω 5Ω ix

5Ω + _ 21 cos 105t V

1 μF

1 μF _ 10.5 cos 105t V +



1 mF 1.6



2A



−j4 Ω

−j3 A

−j4 Ω

Figure P7.60: Circuit for Problem 7.60.

*7.58 Apply nodal analysis in the phasor domain to determine iC (t) in the circuit of Fig. P7.58.

iC

V2



Figure P7.57: Circuit for Problem 7.57.

20 mH

4A

V1

7.61 With Is = 12 120◦ V in the circuit of Fig. P7.61, apply the by-inspection method to develop a node-voltage matrix equation and then use MATLAB or MathScript software to solve for Ix .

20 mH 10 Ω

5Ω + _ 12 cos (400t − 30) V



20 Ω 10 Ω

Figure P7.58: Circuit for Problem 7.58.

7.59 The circuit in Fig. P7.59 contains a supernode between nodes V1 and V2 . Apply the supernode method to determine V1 , V2 , and V3 , and then calculate IC .

Ix

20 Ω

20 Ω Is −j20 Ω

−j10 Ω

Figure P7.61: Circuit for Problem 7.61.

*7.62 Apply nodal analysis to determine IC in the circuit of Fig. P7.62.

“book” — 2015/5/4 — 7:17 — page 452 — #68

452

CHAPTER 7 AC ANALYSIS

8

+

_

IC

2Ω 45o

7.66 Apply the by-inspection method to develop a meshcurrent matrix equation for the circuit in Fig. P7.66 and then use MATLAB or MathScript software to solve for I1, I2, and I3.

2IC

j3 Ω

−j1 Ω

+ V _



7.67 Use any analysis technique of your choice to determine iC (t) in the circuit of Fig. P7.67.

3iC

Figure P7.62: Circuits for Problems 7.62 and 7.63.

7.63 Apply mesh analysis to determine IC in the circuit of Fig. P7.62.

1 mH

20 Ω

5Ω iC

7.64 Apply mesh analysis to determine iL (t) in the circuit of Fig. P7.64. 20 377

10 Ω 104t A

6 cos 2.5

10 Ω

1 μF

mF Figure P7.67: Circuit for Problem 7.67.

iL(t)

+ _ 12 cos 377t (V)

10 377

30 Ω

+ H 4 sin 377t (V) _

*7.68 Determine ix (t) in the circuit of Fig. P7.68, given that υs (t) = 6 cos 5 × 105 t V.

Figure P7.64: Circuit for Problem 7.64.

ix *7.65 Use mesh analysis to obtain an expression for the phasor Vout in the circuit of Fig. P7.65, in terms of Vs and R, given that R = ωL = 1/ωC.

R

20 μF 0.1 Ω 0.1 Ω

+

+ Vs _

R

Vout

L

C

0.2 Ω

20 μF

Figure P7.68: Circuit for Problem 7.68.

_ 7.69 Find Is in the circuit of Fig. P7.69, given that Vs = 8 15◦ V.

Figure P7.65: Circuit for Problem 7.65.



j5 Ω

5Ω + (12 + j6) V _

υs(t) +_

5Ω I1

−j10 Ω

I2

−j10 Ω

Figure P7.66: Circuit for Problem 7.66.

I3

_ + (4 + j3) V

“book” — 2015/5/4 — 7:17 — page 453 — #69

PROBLEMS

453

+ υa _

−j7 Ω 1.5 Ω

j3 Ω

is(t)

_ +

Vs

R2



−j6 Ω

C

R1

L

2.5 Ω Figure P7.72: Circuit for Problem 7.72.

Is Figure P7.69: Circuit for Problem 7.69.

7.73 Find is (t) in the circuit of Fig. P7.73, given that υs = 15 cos(ωt), and:

7.70 Find Z in the circuit of Fig. P7.70, given that Vs = 40 V ◦ and Va = 17.22e−j 132.2 V.

(a) ω = 50 rad/s (b) ω = 75 rad/s (c) ω = 200 rad/s

j4 Ω

Vs

+ _



+ _ Va

−j5 Ω

Z

10 mF is(t) υs(t)

+ _

20 Ω

12 Ω

30 mH

Figure P7.70: Circuit for Problem 7.70.

Figure P7.73: Circuit for Problem 7.73.

*7.71 Find the Th´evenin equivalent at terminals (a, b) for the circuit in Fig. P7.71. The source is Vs = 10 45◦ V. *7.74 Find ia (t) in the circuit of Fig. P7.74, given that is (t) = 18 cos(35t + 75◦ ) A.

Vs

+ _

j6 Ω

4Ω a



b 20 mF

−j5 Ω

24 Ω ia(t)

Figure P7.71: Circuit for Problem 7.71.

7.72 Find ω such that υa (t) and is (t) in the circuit of Fig. P7.72 are in-phase. The element values are R1 = 5 , R2 = 3 , L = 35 mH, and C = 7 mF.

50 Ω

80 mH

is(t)

Figure P7.74: Circuit for Problem 7.74.

30 mH

“book” — 2015/5/4 — 7:17 — page 454 — #70

454

CHAPTER 7 AC ANALYSIS

7.75 Find the value of ω at which υs (t) and is (t) in the circuit of Fig. P7.75 are in phase.

10 Ω

j6 Ω 20 V

30 mH

+ _

is(t) + _

υs(t)

10 A 5 mF

+ _Va

−j4



+ _

−j10 Ω

15 Ω

30 V

Figure P7.78: Circuit for Problem 7.78. Figure P7.75: Circuit for Problem 7.75.

7.79

Find Vo in the circuit of Fig. P7.79.

−j4 Ω

7.76 In the circuit of Fig. P7.76, find the value of ω, given that υs (t) = 20 cos(ωt + 31.4◦ ) V and is (t) = 50 sin(ωt + 80◦ ) A.

is(t)

υs(t)

0.3 Ω



10 V

+ _



_ +

+

−j2 Ω

V0

_

8 mF

+ _

3 mH

Figure P7.79: Circuit for Problem 7.79.

*7.80 The input signal in the op-amp circuit of Fig. P7.80 is given by

Figure P7.76: Circuit for Problem 7.76.

υin (t) = V0 cos ωt. *7.77 Find Ia in the circuit of Fig. P7.77, given that Vs = 10 V and Is = 5 30◦ A.

−j4 Ω

Vs

+ _

Assuming the op amp is operating within its linear range, obtain an expression for υout (t) by applying the phasor-domain technique and then evaluate it for ωRC = 1.





Is

C j4 Ω

Ia

Figure P7.77: Circuit for Problem 7.77.

7.78 Use the superposition principle to solve for Va in the circuit of Fig. P7.78.

+ υin(t) _

R

__ +

+ RL

υout

Figure P7.80: Op-amp circuit for Problem 7.80.

7.81 The input signal in the op-amp circuit of Fig. P7.81 is given by υin (t) = 0.5 cos 2000t V.

“book” — 2015/5/4 — 7:17 — page 455 — #71

PROBLEMS

455

Obtain an expression for υout (t) and then evaluate it for R1 = 2 k , R2 = 10 k , and C = 0.1 μF.

7.84 The signal voltage at the input of a half-wave rectifier circuit is given by υin (t) = A cos(377t + 30◦ ) V. Determine and plot the waveform of υout (t). Calculate the fraction of a full period over which υout = 0 for each of the following values of A (assume VF = 0.7 V):

R2 C R1

_

+ υin(t) _

Section 7-12: Power-Supply Circuits

(a) A = 0.5 V

υout

+

(b) A = 5 V 7.85 A bridge rectifier is driven by a 1 kHz input signal with an amplitude of 10 V. The smoothing filter at the rectifier output uses a 1-μF capacitor in parallel with a load resistor RL . If RD = 5 :

Figure P7.81: Op-amp circuit for Problem 7.81.

(a) What should RL be so that τdn /τup = 2500? 7.82 For υi (t) = V0 cos ωt, obtain an expression for υout (t) in the circuit of Fig. P7.82 and then evaluate it for V0 = 4 V, ω = 400 rad/s, R = 5 k , and C = 2.5 μF.

+ _

υi

(b) How does τdn compare with the period of the rectified waveform? (c) What is the approximate peak value of the output waveform? 7.86 A power supply with the circuit configuration shown in Fig. 7-40 has the following specifications: υs = 24 cos(2π × 103 t + 30◦ ) V, N2 /N1 = 2, C = 0.1 mF, Rs = 50 , RL = 20 k , Rz = 20 , and Vz = 42 V. Determine υout and the peak-to-peak ripple voltage.

υout C

R Section 7-13: Multisim Analysis Figure P7.82: Circuit for Problem 7.82.

*7.83 For υi (t) = V0 cos ωt, obtain an expression for υout (t) in the circuit of Fig. P7.83 and then evaluate it for V0 = 2 V, ω = 377 rad/s, R1 = 2 k , R2 = 10 k , and C = 0.5 μF.

C υi

R1

7.87 Use the Network Analyzer (see Appendix C) in Multisim to determine the equivalent impedance Zeq of the circuit in Fig. P7.87. Using the Network Analyzer, plot Zeq from 1 kHz to 1 MHz and provide a hand calculation demonstrating that the simulated results are correct.

R2

_ +

Figure P7.83: Circuit for Problem 7.83.

180 mH

+

a Zeq

υout

100 kΩ

_

b

Figure P7.87: Circuit for Problem 7.87.

“book” — 2015/5/4 — 7:17 — page 456 — #72

456

CHAPTER 7 AC ANALYSIS

7.88 Use the Network Analyzer (see Appendix C) in Multisim to determine the equivalent impedance Zeq of the circuit in Fig. P7.88. Using the Network Analyzer, plot the real and imaginary parts of Zeq from 100 Hz to 100 kHz and provide a hand calculation demonstrating that the simulated results are correct.

virtual op-amp component, construct the phase-shift oscillator shown in the figure and plot the output from 0 to 1.5 ms in Transient Analysis. Determine the frequency and amplitude of the oscillations as well as the DC offset. Note that you may need to decrease the maximum time step (TMAX) in order to get a clear plot.

a

R4

1 μF 180 μH

R5

Zeq

5 kΩ + _ 1V

1Ω b Figure P7.88: Circuit for Problem 7.88.

R1 C1

R1 C1

R1

R2

R3

5 kΩ C1 5 nF

5 kΩ C2 5 nF

5 kΩ C3 5 nF

υout

Figure P7.90: Circuit for Problem 7.90.

7.89 A 1 V, 100 MHz voltage source υs (t) sends a signal down two transmission lines simultaneously, as depicted by Fig. P7.89. Model this circuit in Multisim with R1 = R2 = 10 , C1 = 7 pF, and C2 = 5 pF and answer the following questions.

R1

1 MΩ

_ +

R1 C1

7.91 Using a Multisim tool or analysis of your choice, find the phase and magnitude of the voltage at each node in the circuit in Fig. P7.91.

R1 C1

+ C1

υout1

_

R2 + υs(t) _

R2 C2

R2 C2

R2 C2

R2 C2

+ C2

υout2

_

Figure P7.89: Circuit for Problem 7.89.

(a) What is the phase shift between υs (t) and the two output nodes, υout1 and υout2 ? (b) What is the amplitude ratio for υout1 /υs and υout2 /υs ? 7.90 Phase-shift circuits have many uses. They can be the fundamental component of an oscillator (a circuit which produces a repetitive electronic signal). The circuit shown in Fig. P7.90 is a phase-shift oscillator. While a detailed analysis is too complex for this text, Multisim allows us to easily create and analyze this circuit. Using the 3-terminal

Potpourri Questions 7.92 Select two from among the touchscreen sensing mechanisms depicted in Fig. TF18-1 of Technology Brief 18. Compare and contrast their advantages and limitations. 7.93 What is the “crystal” in a crystal oscillator? How is it related to piezoelectricity?

“book” — 2015/5/4 — 7:17 — page 457 — #73

PROBLEMS

457

R1 50 Ω 5 cos(2π

+ 103t) V _ V1

C2

R2 C1

3.9 mF 50 Ω

L1

10 mH

R3 50 Ω

3.9 mF L2

10 mH

C3

6.9 mF

R4

100 Ω

Figure P7.91: Circuit for Problem 7.91.

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following prob-lems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .]

(c) Does the circuit seem to “change its personality” with different source frequencies? Explain your answer.

R1

R1 R2 C L

b Figure m7.1 Circuit for Problem m7.1.

m7.2

C

~

R2 b

Figure m7.2 Circuit for Problem m7.2. m7.3 Phase-Shift Circuits: Figure m7.3 shows a phase-shift circuit based on op amps. (a) Write the general expression for the magnitude of Vout with frequency taken as a variable. Hint: View the circuit as the cascade of two standard op-amp circuits. (b) Write the general expression for the phase of Vout with frequency taken as a variable. (c) Set C = 0.1 μF and set all resistors to 1.0 k . Determine the frequency in Hz at which Vout and Vin share the same magnitude. What is the phase shift at this frequency?

C

Equivalent Circuits: For the circuit in Fig. m7.2:

(a) Determine the Th´evenin equivalent circuit at terminals (a, b) using the open-circuit/short-circuit method. Show the Th´evenin impedance as a resistor in series with a single reactive element (capacitor or inductor) and determine the values of all components in the equivalent circuit. The sinusoidal source is Vs = 3 V and f = 500 Hz. R2 = 100 , Component values are: R1 = 90 , C = 1.0 μF, and L = 33 mH. (b) Repeat with the source frequency increased to 1100 Hz.

a

+ υ _ s

m7.1 Impedance Transformations: Determine the equivalent impedance Z looking into terminals (a, b) for the circuit of Fig. m7.1 at the following frequencies: 100 Hz, 500 Hz, 1000 Hz, and 2000 Hz. Report your results in polar form. Use these component values: R1 = 100 , R2 = 90 , C = 1.0 μF, and L = 33 mH.

a

L

R3 R1

υin

+ _

~

_ +

R2

_ +

Figure m7.3 Circuit for Problem m7.3.

υout

“book” — 2015/5/4 — 7:17 — page 458 — #74

458

CHAPTER 7 AC ANALYSIS

m7.4 Introduction to Bode Plots: Determine Va for the circuit of Fig. m7.4 for V1 = 1 V at the following frequencies: 100 Hz, 500 Hz, and 2000 Hz. Use these component values: R1 = 4.7 k , R2 = 3.3 k , R3 = 2.2 k , C1 = 0.047 μF, and C2 = 0.1 μF. For the myDAQ and Multisim portions of this question, use a Bode plotter to capture a representation of Va . The Bode plot provides the magnitude and phase of the gain ratio Va /V1 as a function of frequency. Use cursors to verify the gain and phase of Va from the analytical portion.

R1

V1

~+_

R2

Va C1

R3

Vb C2

m7.6 Arbitrary Sources: Use the myDAQ’s arbitrary waveform generator “ARB” and the AO 0 and AO 1 ports to create the circuit in Fig. m7.6. Then use Multisim to verify your answer (no analytical component to this problem). • V1 is a square wave with LO value: 0, HI value: 1 V, period: 1 ms, and 50% duty cycle. • V2 is a square wave with LO value: 0 V, HI value: 0.5 V, period: 0.5 ms, and 50% duty cycle. (a) From the waveform of the voltage across R4 , determine the four distinct voltage levels (you will need to use the myDAQ’s oscilloscope and Arbitrary Waveform Generator). (b) Using your answer from part (a), find the four corresponding currents flowing through resistor R4 . Verify your answer using Multisim.

Figure m7.4 Circuit for Problem m7.4. m7.5 Frequency Response: An ac circuit may respond differently at different frequencies. Find the peak-to-peak voltage across the 33 k resistor in Fig. m7.5 at each of following frequencies: (a) 1 kHz (b) 2 kHz (c) 20 kHz The amplitude of the ac source is 1 V.

+ R3

V1

22 kΩ

VR4

_

R4 5.6 kΩ

R1

R5

1 kΩ

R2

1 kΩ

1 kΩ V2

Figure m7.6 Circuit for Problem m7.6.

L

+ υs(t)

~

R1

3.3 mH R2

R3 33 kΩ

_

Figure m7.5 Circuit for Problem m7.5.

“book” — 2015/5/4 — 7:19 — page 459 — #1

8

CHAPTER

ac Power Contents Overview, 460 Periodic Waveforms, 460 Average Power, 463 The Electromagnetic Spectrum, 465 Complex Power, 467 The Power Factor, 472 Maximum Power Transfer, 476 Seeing without Light, 477 Measuring Power with Multisim, 482 Summary, 485 Problems, 486

Objectives Learn to: 

Calculate the average and rms value of a periodic waveform.



Determine the complex power, average real power, and reactive power for any complex load with known input voltage or current.



Determine the power factor for a complex load and evaluate the improvement realized by compensating the load through the addition of a shunt capacitor.



Choose the load impedance so as to maximize the transfer of power from the input circuit to the load.



Apply Multisim to measure power.

Rs

Vs

a

CM

RM

c RL

ZSource

+

8-1 8-2 TB20 8-3 8-4 8-5 TB21 8-6

_

ZLoad ZLoad + Match Source

b Matching network

d

LL Load

A matching network is a circuit used to optimize the transfer of ac power between a source and a load. This chapter provides the tools to analyze circuits from the perspective of the total complex power they consume (in their resistors) and store (in their capacitors and inductors).

“book” — 2015/5/4 — 7:19 — page 460 — #2

460

CHAPTER 8 AC POWER

Overview The power absorbed by a resistor R when a current i passes through it is given by p = i 2 R. If i is time varying, we usually designate it i(t), call it the instantaneous current, and call the corresponding power the instantaneous power p(t) = i 2 (t) R

(W).

(8.1)

Usually, we are interested in the average power P consumed by a given circuit—or by a collection of circuits, as in an entire household—and since ac signals are periodic in time with an angular frequency ω and a time period T = 2π/ω, we define Pav to be the average value of p(t) over one (or more) complete period(s). For an ac current given by i(t) = Im cos ωt,

(8.2)

the average power consumed by a resistor R is Pav =

1 2

Im2 R

(W).

(8.3)

This result, which we will derive in Section 8-2, is somewhat intriguing, primarily because Pav is independent of ω and its expression contains a factor of 1/2. The explanations are fairly straightforward and will be covered later. The more important point we wish to make at this time is that had our intent been to discuss ac power in resistive circuits only, the discussion would not have required more than just a few pages and perhaps no more than one or two examples. Instead, we are devoting this entire chapter to ac power because real circuits contain more than just resistors; they contain capacitors and inductors, both of which cannot consume power but can store it and then release it. The current through a resistor is always in phase with the voltage across it. This phase attribute is responsible, in part, for the functional form of the expression for Pav given by Eq. (8.3). The expression, however, generally is not valid when the load circuit contains reactive elements (capacitors and inductors) either alone or in combination with resistive elements. So, for the general case, we need to develop a formulation appropriate for any complex load—from the purely resistive to the purely reactive. That defines one of the objectives of the present chapter. When we transform a circuit from the time domain to the phasor domain, voltages and currents are assigned phasor counterparts, and passive elements become impedances. What about power? Is there a phasor power P, corresponding to p(t)? The answer is: Not exactly. We will introduce a quantity S which we will call complex power, but S is not the phasor counterpart of p(t). In fact, we assign it the symbol S (rather

than P) to avoid the possible misinterpretation that it bears a one-to-one correspondence to p(t). As we will see in Section 8-3, S consists of a real part and an imaginary part with the real part representing the real average power consumed by the circuit and the imaginary part representing the average power stored by the circuit. Towards the end of Chapter 3, we posed the question: When an input circuit is connected to a resistive load, under what condition(s) is the power transferred from the circuit to the load a maximum? Through the application of Th´evenin’s theorem, we demonstrated that the transferred power is a maximum when the load resistance is equal to the Th´evenin resistance of the input circuit. In the present chapter, we pose the question again, but we generalize the load to a complex load ZL = RL + j XL , composed of a resistive part RL and a reactive part XL . In view of the fact that ZL consists of two parts, we should expect the answer to consist of two conditions (not just one) and it does. The details are given in Section 8-5.

8-1

Periodic Waveforms

Even though the focus of this chapter is on the ac power carried by sinusoidally time-varying signals, we will preface our examination by first reviewing some of the important properties shared by all periodic waveforms, including sinusoids. Mathematically, a periodic waveform x(t) with period T satisfies the periodicity property x(t) = x(t + nT )

(8.4)

for any integer value of n. The periodicity property simply states that the waveform of x(t) repeats itself every T seconds. Figure 8-1 displays the waveforms of three typical (and unrelated) periodic functions. In part (a), υ(t) is a sine wave; in (b) i(t) is a sawtooth with a clipped top; and part (c) displays a function given by p(t) = Pm cos2 ωt.

8-1.1

Instantaneous and Average Values

Each of the three waveforms shown in Fig. 8-1 describes the exact variation of its magnitude as a function of time. Consequently, the time function υ(t), for example, is referred to as the instantaneous voltage. Similarly, i(t) is the instantaneous current, and p(t) is the instantaneous power. Often times, however, we may be interested in specifying an attribute of the waveform that conveys useful information about it, and yet it is much simpler to use than the complete waveform. When ac circuits are concerned, two attributes of particular interest are the average value of the waveform and its rootmean-square (rms) value. The latter is introduced in the next

“book” — 2015/5/4 — 7:19 — page 461 — #3

8-1

PERIODIC WAVEFORMS

461 The voltage waveform shown in Fig. 8-1(a) is given by

υ(t) υ(t) = Vm sin (2πt/T)

Vm

T/2

0 −Vm

υ(t) = Vm sin

Vav = 0 T/4

3T/2 T

(8.6)

Application of Eq. (8.5) gives

t

2T

2π t . T

Vav

Positive integrand Negative integrand

(a)

1 = T

T Vm sin 0

Vm = T

i(t)



T − 2π

2π t dt T



 2π t T Vm cos [1 − 1] = 0. (8.7) =−  T 0 2π

Im The fact that the average value of a sine wave is zero is not at all surprising; it is clear from the characteristic symmetry of its waveform that the area under the curve (integrand) during the first half of any cycle is equal (but opposite in polarity) to the area under the curve during the second half of the cycle, so the net sum of the two is exactly zero. In contrast, the lack of symmetry in the waveforms of i(t) and p(t) in Figs. 8-1(b) and (c) (between that part of the waveform above the t axis and the part below it) is an obvious indicator that their average values are not only nonzero but also positive.

Iav 0

T/2

T

3T/2

t

2T

−Im (b) p(t) Pm

p(t) = Pm cos2 (2πt/T) Pav

Pm /2 (c)

T/2

3T/2

T

2T

t

Example 8-1: Average Values

Determine the average values of the waveforms displayed in parts (b) and (c) of Fig. 8-1.

Figure 8-1: Examples of three periodic waveforms.

Solution: During the first half of the first cycle, i(t) is described by a linear ramp of the form subsection, so for the present we pursue only the former. The average value of a periodic function x(t) with period T is given by

Xav

1 = T

T

i(t) = at + b

for 0 ≤ t ≤

T . 2

Its slope is x(t) dt.

(8.5)

a=

0

2Im 4Im = , T /2 T

and its intercept at t = 0 is  We note that Xav is obtained by integrating x(t) over a complete period T and then normalizing the integrated value by dividing it by T . The limits of integration are from 0 to T , but the definition is equally valid for any two limits so long as the upper limit is greater than the lower limit by exactly T (such as from T0 to T0 + T ) or an integer multiple of T . 

b = −Im . Hence,  i(t) =

[(4t/T ) − 1]Im Im

for 0 ≤ t ≤ T /2, for T /2 ≤ t < T .

(8.8)

“book” — 2015/5/4 — 7:19 — page 462 — #4

462

CHAPTER 8 AC POWER

By Eq. (8.5), the average value of i(t) is T 1 i(t) dt Iav = T

8-1.2

For a periodic current waveform i(t) flowing through a resistor R, the average power absorbed by the resistor is

0

⎡ 1 ⎢ = ⎣ T

T /2 0



4t − 1 Im dt + T

T

T /2

Root-Mean-Square (rms) Value

⎤ ⎥ Im Im dt ⎦ = . (8.9) 2

We also can obtain the same result by adding up the areas bounded by one cycle of the waveform of i(t) and then dividing the net total by T , as in

 1 T T T 1 1 Im Iav = − Im × + Im × + I m × = . T 2 4 2 4 2 2 To determine Pav of p(t), we apply Eq. (8.5) to the cos2 function:   T 2π t 1 Pm cos2 dt. Pav = T T 0

The integration is facilitated by applying the trigonometric relation 1 1 cos x = + cos 2x, 2 2

Pav

1 = T

T 0

1 p(t) dt = T

T i 2 (t) R dt.

(8.11)

0

 We would like to introduce a new attribute of i(t), called its effective value, Ieff , defined such that the average power Pav delivered by i(t) to resistor R is equivalent to 2 R.  what a dc current Ieff would deliver to R, namely Ieff That is, 2 R Ieff

= Pav

1 = T

T i 2 (t) R dt.

(8.12)

0

Solving for Ieff gives

2

which leads to the final result

Pm . 2 We should take note for future reference of the fact that the average value of cos2 ωt is 1/2. In fact, it is easy to show that Pav =

1 T



T cos2 0

2πnt + φ1 T

 dt =

1 , 2

and 1 T

(8.10) 

T sin 0

2

2πnt + φ2 T

 dt =

1 , 2

Ieff

   T 1 = i 2 (t) dt T 0

According to Eq. (8.13), Ieff is obtained by taking the square root of the mean (average value) of the square of i(t). The three terms characterizing the operation are coupled together to form root-mean-square (rms for short) and Ieff is relabeled Irms . Even though the idea to define an effective or rms value is introduced in connection with a periodic current waveform, the definition is equally applicable to a periodic voltage waveform as well as to any other periodic waveform. For a periodic waveform x(t), its rms value therefore is defined as

for any values of φ1 and φ2 .  The average values of cos2 (nωt) and sin2 (nωt) are both 1/2 for any integer values of n equal to or greater than 1, irrespective of whether or not their arguments are shifted by constant phase angles, so long as the averaging process is performed over a complete period T = 2π/ω. That is, the average values of cos2 (nωt + φ) and sin2 (nωt + φ) also are 1/2 for any constant value of φ. 

(8.13)

Xrms = Xeff

   T 1 = x 2 (t) dt. T

(8.14)

0

When a multimeter is used to measure an ac voltage waveform, it records the rms value of the voltage. In contrast, when the waveform is displayed on an oscilloscope, the entire waveform is displayed.

“book” — 2015/5/4 — 7:19 — page 463 — #5

8-2 AVERAGE POWER

463

Example 8-2: rms Values

Exercise 8-1: Determine the average and rms values of

Determine the rms values of (a) υ(t) = Vm sin(2π t/T +φ) and (b) i(t) of Fig. 8-1(b).

the waveform υ(t) = 12 + 6 cos 400t V. Answer: Vav = 12 V, Vrms = 12.73 V. (See

)

Solution: (a) Application of Eq. (8.14) to υ(t) gives ⎡ Vrms = ⎣



T

1 T

Vm2 sin2 0

2π t +φ T

Exercise 8-2: Determine the average and rms value of the

⎤1/2



dt ⎦

waveform (8.15)

In view of Eq. (8.10),

Answer: Iav = 0, Irms = 7.48 A. (See

Vm Vrms = √ . 2

(8.16)

Hence:  For any sinusoidal function, its rms value √ is equal to its maximum value (its amplitude) divided by 2.  (b) From Eq. (8.8) of Example 8-1, i(t) is given by  i(t) =

i(t) = 8 cos 377t − 4 sin(377t − 30◦ ) A.

4t T

 − 1 Im

Im

for 0 ≤ t ≤ T2 , for T2 ≤ t < T .

)

8-2 Average Power The circuit configuration shown in Fig. 8-2 consists of an active ac circuit supplying power to a passive load. The load circuit is not restricted in terms of either its architecture or the combination of resistors, capacitors, and inductors it may contain. The instantaneous voltage across the load is υ(t) and the corresponding instantaneous current flowing into it— whose direction is defined in accordance with the passive sign convention—is i(t). Since this is an ac circuit, all of its currents and voltages oscillate sinusoidally at the same angular frequency ω. The general functional forms for υ(t) and i(t) are given by

Its rms value therefore is given by

Irms

⎧ ⎡ ⎤⎫1/2 ⎪ ⎪ 2 T ⎨ 1 T /2 4t ⎬ ⎢ ⎥ 2 2 = , − 1 Im dt + Im dt ⎦ ⎣ ⎪ ⎪ T ⎩T ⎭ T /2

0

which leads to 2Im Irms = √ = 0.82Im . 6

υ(t) = Vm cos(ωt + φυ ),

(8.17a)

i(t) = Im cos(ωt + φi ),

(8.17b)

and

where Vm and Im are the amplitudes of υ(t) and i(t), and φυ and φi are their phase angles, respectively. Our objective is to relate the average power absorbed by the load Pav to the parameters of υ(t) and i(t).

Concept Question 8-1: What is the average value of a sinusoidal waveform? What is its rms value? (See )

Concept Question 8-2: Why is Eq. (8.10) true,

irrespective of the values of φ1 and φ2? Explain in terms of a diagram. (See )

Concept Question 8-3: What does rms stand for and how

does it relate to its definition? (See

)

a i(t) Active ac circuit Source circuit

+ υ(t)

Passive circuit

b

Load circuit

_

Figure 8-2: Passive load circuit connected to an input source at terminals (a, b).

“book” — 2015/5/4 — 7:19 — page 464 — #6

464

CHAPTER 8 AC POWER (b) The waveforms of υ(t) and i(t) are separated from each other by a time shift t, corresponding to the difference in phase angle between them. Given that φυ = +30◦ and φi = −30◦ , υ(t) leads i(t) by 60◦ . A complete period

10 Power p(t) 5

T = 1/f = 1/60 = 16.67 ms

Current i(t) 0

5

10

15 20 Voltage υ(t)

t (ms)

corresponds to a total phase angle of 360◦ . Hence, υ(t) leads i(t) by 

−5

t =

Figure 8-3: Waveforms for a 60 Hz circuit with

υ(t) = 4 cos(377t + 30◦ ) V, i(t) = 3 cos(377t − 30◦ ) A, and p(t) = υ(t) i(t). The waveform of i(t) is shifted by 60◦ behind that of υ(t), and the oscillation frequency of p(t) is twice that of υ(t) or i(t).

The instantaneous power flowing into the load circuit is

cos x cos y =

1 1 cos(x − y) + cos(x + y), 2 2

Pav (8.19)

p(t) can be cast in the form V m Im V m Im cos(φυ − φi ) + cos(2ωt + φυ + φi ). p(t) = 2 2 (8.20) Before proceeding to find the average value of p(t), let us briefly examine the significance of the two terms of Eq. (8.20). The first term is a constant, as it contains no dependence on t, and the second term is sinusoidal, but its angular frequency is 2ω. Thus:  p(t) is the sum of a dc-like term and an ac term that oscillates at a frequency twice that of i(t) and υ(t).  This behavior is evident in the waveforms of υ(t), i(t), and p(t) displayed in Fig. 8-3. The angular frequency ω = 2πf corresponds to f = 60 Hz, and the phase angles were arbitrarily chosen as φυ = 30◦ and φi = −30◦ . The waveform patterns elicit the following observations: (a) The voltage υ(t) oscillates symmetrically relative to the t axis, with a peak-to-peak variation extending from −4 V to +4 V. The current i(t) exhibits a similar pattern between −3 A and +3 A.

 × 16.67 ms = 2.78 ms.

(c) The waveform of the power p(t) is not symmetrical with respect to the t axis. It has a dc shift equal to the first term in Eq. (8.20). Also, it traces twice as many cycles per unit time, in comparison with the waveforms of υ(t) or i(t). Returning to the task at hand, we now apply Eq. (8.5) to the expression of p(t) given by Eq. (8.20) to determine Pav , the average power delivered to the load:

p(t) = υ(t) i(t) = Vm Im cos(ωt + φυ ) cos(ωt + φi ). (8.18) By applying the trigonometric identity

60◦ 360◦

T

1 = T

p(t) dt 0

1 = T

T

Vm Im [cos(φυ − φi ) + cos(2ωt + φυ + φi )] dt. 2

0

(8.21) For any sinusoidal function with ω = 2π/T , it is fairly straightforward to show that for integer values of n equal to or greater than 1 and any constant angle θ

1 T

T cos(nωt + θ) dt = 0 (n = 1, 2, . . . ).

(8.22)

0

Thus, the average value over a period T = 2π/ω of a sinusoidal function of angular frequency ω or integer multiple of ω is zero. In view of Eq. (8.22), the integral of the second term in Eq. (8.21) is zero. Consequently, the expression for Pav simplifies to

Pav =

Vm Im cos(φυ − φi ) 2

(W).

(8.23)

“book” — 2015/5/4 — 7:19 — page 465 — #7

TECHNOLOGY BRIEF 20: THE ELECTROMAGNETIC SPECTRUM

Technology Brief 20 The Electromagnetic Spectrum Electromagnetic Energy The sun’s rays, the signal transmitted by a cell phone, and the radiation emitted by plutonium share a fundamental property: they all carry electromagnetic (EM) energy. It is an interesting and fundamental observation that this energy can be described both as a wave moving through space and as a particle. Neither model alone is sufficient to explain the phenomena we observe in the world around us. This correspondence, called the waveparticle duality, sparked scientific debate as far back as the 1600s, and it was not until the 20th century and the advent of quantum mechanics that this duality was fully incorporated into modern physics. When we treat EM energy as a wave with alternating electric and magnetic fields, we ascribe to the wave a wavelength λ and an oscillation frequency f, whose product defines the velocity of the wave u as u = f λ. If the propagation medium is free space, then u is equal to c, which is the speed of light in vacuum at 3 × 108 m/s. Because of the wave-particle duality, when EM energy is regarded as a particle, each such particle will have the same velocity u as its wave counterpart and will carry energy E whose magnitude is specified by the frequency f through E = hf,

100% Atmospheric transmission

1 mm

wavelength

Figure TF20-1: The electromagnetic spectrum extends over a wide range of wavelengths—from gamma rays to radio waves. The atmosphere is transparent in the microwave and in selected windows in the visible and infrared.

465

where h is Planck’s constant (6.6 × 10−34 J·s). In view of the direct link between E and f, we can refer to an EM particle (also called a photon) either by its energy E or by the frequency f of its wave counterpart. The higher the frequency is, the higher is the energy carried by a photon, but also the shorter is its wavelength λ.

The Spectrum In terms of the wavelength λ, the EM spectrum extends across many orders of magnitude (Fig. TF20-1), from the radio region on one end to the gamma-ray region on the other. The degree to which an EM wave is absorbed or scattered as it travels through a medium depends on the types of constituents present in that medium and their sizes relative to λ of the wave. For Earth’s atmosphere, the composition and relative distributions of its gases are responsible for the near total opacity of the atmosphere to EM waves across most of the EM spectrum, except for narrow “windows” in the visible, infrared, and radio spectral regions (Fig. TF20-1). It is precisely because EM waves with these wavelengths can propagate well through the atmosphere that human sight, thermal infrared imaging, and radio communication are possible through the air. 1. Cosmic Rays: Emitted by the decay of the nuclei of unstable elements and by cosmic, high-energy sources in the universe, cosmic rays—which include gamma, beta, and alpha radiation—are highly energetic particles that can be dangerous to organisms and destructive to matter. Earth emits gamma rays of its own, but at very weak levels. 2. X-Rays: Slightly lower energy radiation falls into the X-ray region; this radiation is energetic enough to be dangerous to organisms in large doses, but small doses are safe. More importantly, their relatively high energy allows them to traverse much farther into solid objects than lower frequency radiation (such as visible light). This phenomenon allows for modern medical radiology, in which X-rays are used to measure the opacity of the medium between the X-ray source and the detector or film. Thankfully, Earth’s atmosphere efficiently screens the surface from high-energy radiation, such as cosmic rays and X-rays. 3. Ultraviolet Rays: The atmosphere is only partially opaque to ultraviolet (UV) waves, which border the visible

“book” — 2015/5/4 — 7:19 — page 466 — #8

466

TECHNOLOGY BRIEF 20: THE ELECTROMAGNETIC SPECTRUM

Table TT20-1: Some examples of radio frequency communication channels and their frequency bands.

Communicaon modality Medium wave AM radio (US)

Band name MF

FM radio (US)

VHF

GPS L1 and L2

UHF

802.11g wifi

ISM

Bluetooth® 802.15.4 – ZigBee (US) 802.15.4 – ZigBee (Asia)

ISM UHF UHF

spectrum on the short-wavelength side. UV radiation is both useful in modern technology and potentially harmful to living things in high doses. Among its many uses, UV radiation is used routinely in electronic fabrication technology for erasing programmable memory chips, polymer processing, and even as a curing ink and adhesive. While UV’s potential danger to human skin is well recognized, it is for the same reasons that UV lamps are used to sterilize hospital and laboratory equipment. 4. Visible Light Rays: The wavelength range of visible light extends from about 380 nm (violet color) to 740 nm (red/brown color), although the exact range varies from one human to another. Some species can see well into the infrared (IR) or the UV, so the definition of visible is completely anthropocentric. It is no coincidence that evolution led to the development of sight organs that are sensitive to precisely that part of the spectrum where atmospheric absorption is very low. In the visible spectrum, blue light is more susceptible to scattering by atmospheric particles than the longer wavelengths, which is why the sky appears blue to us. 5. Infrared Rays: The infrared (IR) region, straddled in between the visible spectrum and the radio region, is particularly useful for thermal applications. When an object is heated, the added energy increases the vibrations of its molecules. These molecular vibrations, in turn, release electromagnetic radiation at many frequencies. Within the range of our thermal environment, the peak of the radiated spectrum is in the IR region. This feature has led to the development of IR detectors and cameras for both civilian and military thermal-imaging

Frequencies 520 – 1610 kHz, broken into 10 kHz channels 88 – 108 MHz, broken into 100 – 200 kHz channels 1575.42 MHz (L1) and 1227.60 MHz (L2) 2.4 – 2.5 GHz, broken into 13 overlapping 22 MHz channels 2400 – 2483 MHz, broken into 1 MHz channels 902 – 928 MHz, broken into 30 channels 2.4 GHz, broken into 16 channels

applications. Nightvision systems use IR detector arrays to image a scene when the intensity of visible-wavelength light is insufficient for standard cameras. This is because material objects emit IR energy even in pitch-black darkness. Conversely, IR energy can be used to heat an object, because a good radiator of IR is also a good absorber. Additionally, IR beams are used extensively in short-distance communication, such as in the remote control of most modern TV sets and garage door openers.

6. Radio Waves: The frequency range of the radio spectrum extends from essentially dc (or zero frequency) to f = 1 THz = 1012 Hz. It is subdivided into many bands with formal designations (Fig. TF20-1) such as VHF (30 to 300 MHz) and UHF (300 to 3000 MHz), and some of those bands combine together to form bands commonly known by historic designations, such as the microwave band (300 MHz to 30 GHz). All major freespace communication systems operate at frequencies in the radio region, including wireless local area networks (LANs), cell phones, satellite communication, and television and radio transmissions (Table TT20-1). Because the radio spectrum is used so heavily, spectrum allocation is controlled (often sold) by various national and international agencies that set standards for what types of devices are permitted to operate, within what frequency bands, and at what maximum-power transmission levels. Cell phones, for example, are allowed to transmit and receive in the 2.11 to 2.2 GHz band and in the 1.885 to 2.025 GHz band. Radio waves are the range where the classic antenna-to-antenna transmission occurs.

“book” — 2015/5/4 — 7:19 — page 467 — #9

8-3

COMPLEX POWER

467

This is the dc term in Eq. (8.20). According to Eq. (8.16), the rms value of a sinusoidal√voltage waveform is related to its amplitude by Vrms = Vm / 2, and a similar relationship holds for i(t). Hence,

Exercise 8-3: The voltage across and current through a

certain load are given by υ(t) = 8 cos(754t − 30◦ ) V and

Pav = Vrms Irms cos(φυ − φi )

(W).

i(t) = 0.2 sin 754t A.

(8.24)

The quantity (φυ − φi ) is called the power factor angle and plays a critical role with respect to Pav . For a purely resistive load R, υ(t) and i(t) are in phase, which means that φυ = φi . Consequently,

V2 Pav = Vrms Irms = rms . R (purely resistive load)

(8.25)

What is the average power consumed by the load, and by how far in time is i(t) shifted relative to υ(t)? Answer: Pav = 0.4 W; t = 1.39 ms. (See

8-3

)

Complex Power

The correspondence between the instantaneous voltage υ(t) and instantaneous current i(t) and their respective phasors (V and I) is embodied by the relationships υ(t) = Vm cos(ωt + φυ )

V = Vm ej φυ

(8.27a)

i(t) = Im cos(ωt + φi )

I = Im ej φi .

(8.27b)

and The average power supplied to the load is exactly what we would expect for a resistor at dc, except that in the ac case, we substitute rms for dc values. For a purely reactive load (capacitors and/or inductors, with no resistors), we established in Section 7-7 that (φυ − φi ) = ±90◦ , with the (+) sign corresponding to an inductive load (because υL leads iL by 90◦ ) and the (−) sign corresponding to a capacitive load (υC lags iC by 90◦ ). In either case, Pav = Vrms Irms cos 90◦ = 0.

(8.26)

(purely reactive load)

In the time domain, in general it is not possible to combine all of the elements of a passive load circuit into a single equivalent element, but it is possible to do so in the phasor domain. A passive ac circuit always can be represented by an equivalent impedance Z, as shown in Fig. 8-4, and it has to satisfy the condition Z=

Vm j (φυ −φi ) V = e I Im

( ),

(8.28a)

where V and I are the phasor voltage and current at its input terminals. Since in general Z = |Z| ej φz ,

 A purely reactive load can store power and then release it, but the net average power it absorbs is zero. 

Concept Question 8-4: How is the rms value related to

the amplitude of a sinusoidal signal? (See

)

Active ac circuit (phasor domain)

I

+ V

_

Z

b Source circuit

Concept Question 8-5: How much average power is

consumed by a reactive load? Explain. (See

a

)

Load circuit

Figure 8-4: Source circuit connected to an impedance Z of a load circuit.

“book” — 2015/5/4 — 7:19 — page 468 — #10

468

CHAPTER 8 AC ANALYSIS

where φz is the phase angle of Z, it follows that Vm |Z| = , Im

Hence, S = Pav + j Q

φz = φυ − φi .

(8.28b)

The complex power S is a phasor quantity defined in terms of V and I, but it is not simply the product of V and I. The definition of S is constructed such that the real part of S is exactly equal to Pav , the real average power absorbed by the load Z. To that end, S is defined as S=

1 VI∗ 2

(VA),

(8.29)

where I∗ is the complex conjugate of I, realized by replacing j with −j everywhere in I. Upon inserting the expressions for V and I given by Eqs. (8.27a and b) into Eq. (8.29) (after replacing j φi with −j φi so as to convert I to I∗ ), we obtain the result 1 (Vm ej φυ )(Im e−j φi ) 2 1 = Vm Im ej (φυ −φi ) 2 1 1 = Vm Im cos(φυ − φi ) + j Vm Im sin(φυ − φi ). (8.30) 2 2

S=

For the sake of consistency, we introduce the rms phasor voltage and current as V Vm Vrms = √ = √ ej φυ 2 2

(8.31a)

(VA),

(8.35)

and conversely, Pav = Re[S]

(average absorbed power)

(8.36a)

Q = Im[S]

(peak exchanged power).

(8.36b)

and

 Whereas Pav represents real dissipated power, Q represents the peak amount of power exchanged (back and forth) between the source circuit and the load circuit.  During a single oscillation cycle of duration T : Pav T = energy dissipated in the load, QT = energy transferred to the load and then returned to the source. The three quantities—S, Pav , and Q—are each a product of a voltage and a current and therefore should be measured in watts.  However, to help distinguish between them, only Pav retains the unit of watt, and the other two have been assigned artificially different units. S has been given the unit volt-ampere (VA) and Q the unit volt-ampere reactive (VAR). 

and I Im Irms = √ = √ ej φi , 2 2

(8.31b)

and we rewrite Eqs. (8.29) and (8.30) in terms of rms quantities as ∗ S = Vrms Irms

(VA),

(8.32)

8-3.1

Complex Power for a Load

So far, we have expressed S in terms of V and I, but V and I are linked to one another through the impedance of the load circuit Z (Fig. 8-4). In general, Z has a real, resistive component R, and an imaginary, reactive component X: Z = R + j X.

and S = Vrms Irms cos(φυ − φi ) + j Vrms Irms sin(φυ − φi ). (8.33) We note that the real part of S (first term) is equal to the expression for Pav given by Eq. (8.24). The second term is called the reactive power Q: Q = Vrms Irms sin(φυ − φi )

(VAR).

(8.34)

We should recall from Chapter 7 that the reactive component is inductive if X > 0 and capacitive if X < 0. In terms of Z, V = Z I,

(8.37)

and the expression for S given by Eq. (8.29) becomes S=

1 1 1 2 (R + j X), (8.38) V I∗ = Z I I∗ = |I|2 Z = Irms 2 2 2

“book” — 2015/5/4 — 7:19 — page 469 — #11

8-3

COMPLEX POWER

469

From this, we deduce that

Pav = Re[S] =

Im

1 2 2 |I| R = Irms R 2

(W)

S

Q

(8.39a)

ϕυ − ϕi

and

2 Pav = Irms R

1 2 Q = Im[S] = |I|2 X = Irms X 2

(VAR).

Re

Pav (X > 0)

2 Q = Irms X

(8.39b)

(a) Inductive load The relationship between S and its components Pav and Q is illustrated graphically in Fig. 8-5(a) for an impedance with an inductive component (X > 0). A similar illustration is contained in Fig. 8-5(b) for an impedance with a capacitive component (X < 0). The vector S lies in quadrant 1 (0 < (φυ − φi ) ≤ 90◦ ) if X is inductive and in quadrant 4 (−90◦ ≤ (φυ − φi ) < 0) if X is capacitive. (If S were to lie in quadrants 2 or 3, Pav would be negative, indicating that the load is actually a source supplying power, not consuming it.)

8-3.2

Im 2 Pav = Irms R 2 Q = Irms X

ϕυ − ϕi

Q

Conservation of Complex Power

Re (X < 0)

S (b) Capacitive load

In a circuit containing n elements, energy conservation requires that the sum of the complex powers associated with all n elements be equal to zero: n 

Pav

Si = 0.

Figure 8-5: Complex power S lies in quadrant 1 for an inductive load and in quadrant 4 for a capacitive load.

Example 8-3: RL Load

i=1

Since Si is complex, it follows that both the real and imaginary components of the sum have to individually be equal to zero, which, in view of Eq. (8.35), leads to n  i=1

Pavi = 0,

n 

Qi = 0.

(8.40)

i=1

An input circuit consisting of a source υs = 10 cos 105 t V in series with a source resistance Rs = 100 is connected to an RL load circuit, as shown in Fig. 8-6(a). If R = 300 and L = 3 mH, determine: I, S into the RL load and φυ of the voltage across the load. Solution: From the expression for υs , we deduce that Vs = 10 V and ω = 105 rad/s. Hence, the load impedance is Z = R + j ωL = 300 + j 105 × 3 × 10−3 = (300 + j 300) .

Keeping in mind that Pavi has a positive (+) value if the ith element is a resistor and a negative (−) value if it is a generator of power, the first summation in Eq. (8.40) states that the power consumed by the resistors is equal to the (real) power generated by the sources in the circuit. Similarly, the summation over Qi states that there is no net exchange of reactive power between the sources and the reactive elements in the circuit.

The phasor current I corresponding to i(t) is given by I=

10 Vs = Rs + Z 100 + 300 + j 300 10 ◦ = = 20e−j 36.87 mA. 400 + j 300

“book” — 2015/5/4 — 7:19 — page 470 — #12

470

CHAPTER 8 AC ANALYSIS

Rs

a

and their combination specifies S as

i



S = 60 + j 60 = 84.85ej 45 mVA.

+ R υs

+ −

~

υ L

According to Eq. (8.30), the phase angle of S is equal to φυ −φi . Hence, 45◦ = φυ − (−36.87◦ ),

_ b

Input circuit

which yields

Load

(a) Circuit

φυ = 8.13◦ . We also can determine V independently by applying voltage division to the circuit,

Im V=

S

Q = 60 mVAR 45o

Pav = 60 mW

Re

10(300 + j 300) Vs Z ◦ = = 8.48ej 8.13 V, Rs + Z 400 + j 300

which confirms the value we found earlier for φυ . Figure 8-6(b) and (c) provide graphical renditions of S, V, and I in the complex plane. Example 8-4: Capacitive Load

(b) S in complex plane Im ϕυ − ϕi = 45o V ϕi =

ϕυ = 8.13o

−36.87o

Re

The current source is (t) = 20 cos(103 t + 30◦ ) mA and associated shunt resistance Rs = 400 , as shown in Fig. 8-7(a), provide ac power to the load circuit to the right of terminals (a, b). If R1 = 200 , R2 = 2 k , and C = 1 μF, determine: (a) I, V, S, Pav , and Q for the entire load circuit (to the right of terminals (a, b)), (b) SC for the capacitor alone, and (c) Ss for the current source. Solution: (a) In the phasor domain, ◦

I (V and I have different scales)

Is = 20ej 30 mA and ZC =

(c) V and I in complex plane Figure 8-6: Example 8-3.

−j −j = −j 1000 , = 3 ωC 10 × 10−6

and the impedance Z of the load circuit is Z = R1 + R2  ZC

Given that Im = 20 mA and R = X = 300 , 2 Pav = Irms R=

(20 × 10−3 )2 Im2 R = × 300 = 60 mW 2 2

and 2 Q = Irms X=

(20 × 10−3 )2 × 300 = 60 mVAR, 2

= 200 +

2000 × (−j 1000) = (600 − j 800) . 2000 − j 1000

Current division in the phasor-domain circuit of Fig. 8-7(b) yields ◦

I=

I s Rs 20 × 10−3 ej 30 × 400 ◦ = = 6.25ej 68.66 mA, Rs + Z 400 + (600 − j 800)

“book” — 2015/5/4 — 7:19 — page 471 — #13

8-3

COMPLEX POWER

a

471

i

and the corresponding voltage VC across the capacitor is

R1



VC = IC ZC = 5.59ej 95.23 × 10−3 × (−j 1000)

+ is

Rs

123



R2

υ

_

= 5.59ej 5.23 V,

C

where we used the identity ◦

−j = e−j 90 .

123

b

Input

The complex power associated with the capacitor is 1 1 ◦ ◦ SC = VC IC∗ = 5.59ej 5.23 × 5.59 × 10−3 e−j 95.23 2 2

Load

(a) Time domain



a

I

200 Ω

= 15.62e−j 90 = 0 − j 15.62 mVA.

IC

+ Is

400 Ω V

2000 Ω

_

−j1000 Ω

123

b

Z = (600 −j800) Ω (b) Phasor domain Figure 8-7: Circuit for Example 8-4.

and the phasor voltage at terminals (a, b) is



= −62.5e−j 14.47 mVA ◦

V = IZ = 6.25×10−3 ej 68.66 ×(600−j 800) = 6.25ej 15.53 V. Given I and V, the complex power S is S=

As expected, the real part of SC (representing the amount of power dissipated in the capacitor) is zero, and the imaginary part is exactly equal to Q of the overall load circuit (the capacitor is the only element in the load circuit capable of exchanging power back and forth with the input circuit). (c) Recall that for any device, S represents the complex power transferred into the device, and it is defined such that the current direction through the device is from the (+) terminal to the (−) terminal of the voltage across it. For the current source Is , it flows through itself from the (−) terminal of V to the (+) terminal of V, in exact opposition to the definition of S. Hence, 1 1 ◦ ◦ Ss = − VIs∗ = − × 6.25ej 15.53 × 20e−j 30 × 10−3 2 2

1 1 ◦ ◦ VI∗ = × 6.25ej 15.53 × 6.25 × 10−3 e−j 68.66 2 2 ◦

= 19.53e−j 53.13 mVA,

= −62.5 cos(−14.47◦ ) − j 62.5 sin(−14.47◦ ) = (−60.52 + j 15.62) mVA. The real part of Ss represents the real average power generated by Is and is equal in magnitude to the average power dissipated in the three resistors in the circuit. The imaginary part of Ss is equal in magnitude and opposite in sign to SC . Concept Question 8-6: What are the two components

with real and imaginary components given by Pav = Re[S] = 19.53 × 10−3 cos(−53.13◦ ) = 11.72 mW and

of the complex power S, what type of power do they represent, and what units are assigned to them? (See ) Concept Question 8-7: If S lies in quadrant 2 in the

Q = Im[S] = 19.53 × 10

−3



sin(−53.13 ) = −15.62 mVAR.

complex plane, what does that tell you about the load? (See )

(b) The phasor current IC flowing through C is related to I by ◦

IC =

R2 I 2000 × 6.25 × 10−3 ej 68.66 = R2 + Z C 2000 − j 1000 ◦

= 5.59ej 95.23

(mA),

Exercise 8-4: The current flowing into a load is given by i(t) = 2 cos 2500t A. If the load is known to consist of a series of two passive elements, and S = (10 − j 8) VA, determine the identities of the elements and their values. Answer: R = 5 , C = 100 μF. (See

)

“book” — 2015/5/4 — 7:19 — page 472 — #14

472

CHAPTER 8 AC ANALYSIS

Table 8-1: Summary of power-related quantities.

Time Domain a Input circuit

Phasor Domain

i

υ

I

a Load circuit

Input circuit

Z = R + jX = |Z|e jϕz

V

b

b V = Vm ej φυ I = Im ej φi

υ(t) = Vm cos(ωt + φυ ) i(t) = Im cos(ωt + √ φi ) Vrms = Vm /√2 Irms = Im / 2

Vrms = Vrms ej φυ Irms = Irms ej φi

Complex Power Real Average Power Pav = Re [S] = Vrms Irms cos(φυ − φi ) 2 R = V 2 R/|Z|2 = Irms rms Apparent Power  2 + Q2 S = |S| = Pav = Vrms Irms 2 |Z| = V 2 /|Z| = Irms rms

∗ = P + jQ S = 21 VI∗ = Vrms Irms av

S = Sej (φυ −φi ) = Sej φz φz = φυ − φi

Power Factor Pav pf = S = cos(φυ − φi ) = cos φz

a

8-4 The Power Factor Several power-related terms were introduced in the preceding two sections, including the complex power S, the real average power Pav , and the reactive power Q. We plan to introduce two additional terms in this section, so lest this apparent profusion of terms contribute to any possible confusion, we have prepared a summary of all relevant terms and expressions in the form of Table 8-1. This is intended to provide the reader easy access to and greater clarity about the interrelationships among the various power quantities. In terms of the complex quantities V and I (representing the phasor voltage across a load circuit and the associated current into it) the complex power S transferred to the load circuit (Fig. 8-8) is given by S = Pav + j Q,

Reactive Power Q = Im [S] = Vrms Irms sin(φυ − φi ) 2 X = V 2 X/|Z|2 = Irms rms

Electrical power source

I R Z = R + jωL

V L b

Inductive load

(a) a Electrical power source

(8.41)

I R Z=R−

V C

j ωC

b Capacitive load (b)

with Pav = Vrms Irms cos(φυ − φi )

(8.42a)

Figure 8-8: Inductive and capacitive loads connected to an electrical source.

“book” — 2015/5/4 — 7:19 — page 473 — #15

8-4 THE POWER FACTOR

473

Table 8-2: Power factor leading and lagging relationships for a load Z = R + j X. Load Type Purely Resistive (X = 0)

φz = φυ − φi

I-V Relationship

pf

φz = 0

I in phase with V

1

Inductive (X > 0)

0 < φz ≤ 90◦

Purely Inductive (X > 0 and R = 0)

φz = 90◦

Capacitive (X < 0)

−90◦ ≤ φz < 0

Purely Capacitive (X < 0 and R = 0)

φz = −90◦

and

I lags V

lagging

I lags V by 90◦

lagging

I leads V

leading

I leads V by 90◦

leading

Capacitive load Q = Vrms Irms sin(φυ − φi ).

(8.42b)

For reasons that we will discuss in the next subsection, the magnitude of S is called the apparent power S, and it is given by  2 + Q2 = V S = |S| = Pav rms Irms ,

(8.43)

and the ratio of Pav to S is called the power factor pf, and is given by pf =

Pav = cos(φυ − φi ). S

(8.44)

The argument of the cosine (φυ − φi ) is called the power factor angle. Per Eq. (8.28b), this angle is equal to the phase angle of the load impedance φz : φz = φυ − φi .

(8.45)

In view of Eq. (8.45), the expression for the power factor can be rewritten as pf = cos φz .

(8.46)

Inductive load An inductive load, such as a series RL circuit, has an impedance Zind = R + j ωL.

(8.47)

As both components of Zind are positive quantities, φz is positive. Since R cannot be negative, the range of φz is 0 ≤ φz ≤ 90◦ with 0◦ corresponding to a purely resistive load and 90◦ corresponding to a purely inductive load.

The equivalent circuit of a capacitive load is a series RC circuit with j . (8.48) Zcap = R − ωC Consequently, φz is negative and its range is −90◦ ≤ φz ≤ 0, with −90◦ corresponding to a purely capacitive load. Because cos(−θ) = cos θ for any angle θ between −90◦ and +90◦ , the power factor (Eq. (8.46)) is insensitive to the sign of φz , and therefore, it cannot differentiate between an inductive load and a capacitive load. To qualify pf with such information:  The load is said to have a leading pf or a lagging pf, depending on whether the current I leads or lags the voltage V (see Table 8-2). 

8-4.1

Power Factor Significance

Most industrial loads involve the use of large motors or other inductive machinery that require the supply of tens of kilowatts of power, typically at 440 V rms. Household appliances (such as refrigerators and air conditioners) also contain inductive coils, and most are designed to operate at either 110 V rms or 220 V rms. Thus, most loads to which an electrical source has to supply power have an RL equivalent circuit of the type shown in Fig. 8-8(a). From the perspective of an energy supplier (such as the electric power company) the load has two important attributes: S and Pav . The amount of power the company has to supply is S, but it can charge for only Pav , because Pav is the only real power consumed by the load. The company appears to supply S—hence, the name apparent power—but it gets paid for a fraction of that, and the power factor is that fraction. For two loads—one purely resistive with Z1 = R and the second

“book” — 2015/5/4 — 7:19 — page 474 — #16

474

CHAPTER 8 AC ANALYSIS

inductive with Z2 = R + j ωL—with both requiring the same voltage V and consuming the same power Pav , the inductive load will require the transmission of a larger current to it than would the purely resistive load. This point is demonstrated numerically through Example 8-5.

Is

IL R

Generator circuit

VL

ZL L

Example 8-5: ac Motor

Inductive load The equivalent circuit of a dishwasher motor is characterized by an impedance Z = (20 + j 20) . The household voltage is 110 V rms. Determine: (a) pf, S, and Pav and (b) the current that the electric company would have supplied to the motor had it been purely resistive and consumed the same amount of power. Solution: (a) We will treat the phase of the voltage as our reference by setting it arbitrarily equal to zero. Thus, Vrms = Vrms 0◦ = 110 V.

(a) Uncompensated load Is

IL IC

Generator circuit

C

R VL

ZL L

123

(8.49)

Compensated load This is justified by the fact that none of the quantities of interest require knowledge of the values of φυ and φi individually; it is the difference (φυ − φi ) that counts. The corresponding current is Irms =

Vrms 110 110 = 3.9 −45◦ A, = = √ ◦ Z 20 + j 20 20 2 ej 45

from which we deduce that Irms = 3.9 A and φz = 45◦ . The quantities of interest are then given by S = Vrms Irms = 110 × 3.9 = 427.8 VA, Pav = S cos φz = 429 cos 45◦ = 302.5 W, and pf =

Pav = 0.707. S

(8.50)

(b) A purely resistive load that consumes 302.5 W at 110 V rms must have a current of Irms =

Pav 302.5 = = 2.75 A. Vrms 110

(8.51)

For the same amount of consumed power, the power supplier has to provide 3.9 A to an inductive load with a power factor of 0.707, compared with only 2.75 A to a purely resistive load with pf = 1.

(b) Compensated load Figure 8-9: Adding a shunt capacitor across an inductive load reduces the current supplied by the generator.

8-4.2

Power Factor Compensation

Raising the power factor of an inductive load (such as an electric drill or a compressor) is highly desirable, not only for the energy supplier but also ultimately for its customers as well. Redesigning the load circuit itself to raise its power factor to a value closer to 1, however, may not be practical, primarily because its motor or other inductive components were presumably selected to meet certain operational specifications that may be incompatible with a higher power factor. This problem of partial incompatibility raises the following question: can we raise the pf of a load (as seen by the generator circuit) while keeping it the same as far as the inductive load itself is concerned? The answer is yes, and the solution is fairly straightforward: it entails adding a shunt capacitor across the inductive load, as shown in Fig. 8-9(b). Without the capacitor (Fig. 8-9(a)), the inductor load requires a voltage VL across it and a current IL through it. The source current Is is equal to IL . The presence of the shunt capacitor does not change VL , and by virtue of the load impedance ZL , the current IL = VL /ZL also remains unchanged. In other words, the capacitor exercises no influence on the inductive load, but it does change the overall load circuit as far as the generator is concerned. The new load

“book” — 2015/5/4 — 7:19 — page 475 — #17

8-4 THE POWER FACTOR

475 magnitude) is shorter than the length it was before adding the capacitor. In terms of the power factor,

Im IC



VL ϕnew Is

ϕZL

pf =

Re

I L = Is (a) Phasor currents Im QL

ϕZL

Re

PL

Q = QL + QC

(b) Uncompensated load

(8.53)

(8.54)

is smaller than QL alone, thereby reducing the phase angle from φzL to φnew , where

Im QL SL ϕnew QC

for the RL circuit alone, for the compensated circuit,

where φnew is the phase angle between Is and VL in the compensated load circuit. Another approach to demonstrate how the addition of the capacitor improves the power factor is by comparing the power factor triangle of the RL circuit alone with that of the compensated load circuit that includes the capacitor. The two triangles are diagrammed in parts (b) and (c) of Fig. 8-10, in which PL and QL represent the consumed and reactive powers associated with the RL load, and QC is associated with the capacitor C. The capacitor introduces reactive power QC , and since QC is negative, the net sum

IC

SL

cos φzL cos φnew

Q = QL + QC

PL

φnew = tan−1

Re

Q PL

 .

(8.55)

Example 8-6: pf Compensation

(c) Compensated load Figure 8-10: Comparison of source currents and power factor triangles for the compensated and uncompensated circuits.

circuit—which we will call the compensated load circuit— consists of the parallel combination of C and the original RL circuit. Because of the new current IC , the source current becomes Is = IL + IC .



(8.52)

Had C and the RL load been purely resistive, both IC and IL would have been real and of the same sign, resulting in a larger source current rather than smaller. Fortunately, IC and IL are phasor quantities, and their imaginary components have opposite polarities (actually, IC is purely imaginary). With VL chosen to serve as the phase reference, Fig. 8-10(a) illustrates how the vector sum of IL (the current into the RL circuit) and IC leads to a vector Is , whose length (or equivalently, its

A 60 Hz electric generator supplies a 220 V rms to a load that consumes 200 kW at pf = 0.8 lagging. By adding a shunt capacitor C, the power factor of the overall circuit was improved to 0.95 lagging. Determine the value of C. Solution: A power factor of 0.8 corresponds to a phase angle φzL given by φzL = cos−1 (pf1 ) = cos−1 (0.8) = 36.87◦ . The values of SL and QL for the load alone are SL =

PL 200 × 103 = = 250 kVA pf1 0.8

and QL = SL sin φzL = 250 sin 36.87◦ = 150 kVAR. The associated power triangle is shown in Fig. 8-11(a).

“book” — 2015/5/4 — 7:19 — page 476 — #18

476

CHAPTER 8 AC ANALYSIS

Im

Concept Question 8-8: Why is the power factor of a

household appliance significant to an electric utility company? (See )

VA

k 50

SL

QL = 150 kVAR

=2

pf1 = 0.8

36.87o

Concept Question 8-9: What is pf compensation, and

Re

PL = 200 kW (a)

ZL = (50 + j 50) . (a) What is the value of the power factor of ZL . (b) What will be the new power factor if a 1 capacitance C = 12π mF is added in parallel with the RL load?

pf2 = 0.95 Q = 65.72 kVAR

18.19o

Re

PL = 200 kW

Figure 8-11: Power triangles for Example 8-6.

Addition of the capacitor changes the power factor to pf2 = 0.95, with a corresponding angle as φnew = cos

Answer: (a) pf1 = 0.707, (b) pf2 = 1. (See

8-5

(b)

−1

)

Exercise 8-5: At 60 Hz, the impedance of a RL load is

Im SL

why is it used? (See

−1

(pf2 ) = cos

)

Maximum Power Transfer

Consider the network configuration shown in Fig. 8-12 in which an ac source circuit is represented by its Th´evenin equivalent circuit, composed of a phasor voltage Vs and a source impedance Zs = Rs + j Xs .

(8.56)

Similarly, the load is represented by its impedance ZL with



(0.95) = 18.19 .

ZL = RL + j XL .

(8.57)

The consumed power PL does not change, but from Fig. 8-11(b), the new reactive power is now

a IL

Q = 200 tan φnew = 200 tan 18.19◦ = 65.72 kVAR. Using the value of QL we determined earlier, the reactive power introduced by the capacitor is

Active ac circuit

QC = Q − QL = (65.74 − 150) = −84.26 kVAR.

Source circuit

VL

Passive circuit

b

Load circuit

a

IL

With ZC = 1/j ωC, the complex power of C is ∗ SC = VLrms IC = VLrms rms

∗ VL rms

Z∗C

Zs = Rs + jXs

= −j |VLrms |2 ωC.

Hence, PC = 0, and

Vs QC = −|VLrms |2 ωC.

Solving for C gives 84.26 × 103 −QC = = 4.62 mF. C= 2 2πf Vrms 2π × 60 × (220)2

+ _

VL

ZL = RL + jXL

b Figure 8-12: Replacing the source and load circuits with their respective Th´evenin equivalents.

“book” — 2015/5/4 — 7:19 — page 477 — #19

TECHNOLOGY BRIEF 21: SEEING WITHOUT LIGHT

477

Technology Brief 21 Seeing without Light When we think of optical technology, we most often think of visible light—the wavelengths we can see, that our eyes are sensitive to. These wavelengths range from about 750 nm (red) to 400 nm (violet), as shown in Fig. TF21-1. From f = c/λ, where c = 3 × 108 m/s and λ is the wavelength in meters, the corresponding frequency range extends from 400 THz (1 THz = 1012 Hz) for red light to 750 THz for violet light. Our eyes are insensitive to electromagnetic waves whose frequencies are outside this range, but we can build sensors that are. Infrared (IR) frequencies (those below the visible spectrum) can be used for thermal imaging (sensing heat) and night vision (seeing in the dark). Ultraviolet (UV) frequencies (those above the visible spectrum) can be used for dermal (skin) imaging as well as numerous surface treatments (see Technology Brief 5 on LEDs).

Thermal — Infrared (IR) Imaging Figure TF21-2: A night-vision image taken with military-

Night-vision imaging is used for a wide variety of applications including imaging people for security and rescue (as seen in Figs. TF21-2 and TF21-3). Helicopters can fly over large regions, locating people and animals from their IR signatures. Firefighters can use IR goggles to see through smoke and find victims.Thermal imaging is

grade goggles.

also used for medical applications (inflammation warms injured body parts) including those involving animals and small children who cannot tell you “where it hurts” and

Visible light

Infrared

Ultraviolet

Near

Near A

B C

Far (vacuum)

nm

nm

nm

nm

nm

nm

nm

nm

nm 0 20 nm 0 28 m n

0

0

32

40

0

45

0

50

0

55

0

60

0

65

0

70

μm

0

75

4

μm

m m

1.

3

1

Figure TF21-1: Spectrum of visible light and its two neighbors, the infrared and the ultraviolet.

“book” — 2015/5/4 — 7:19 — page 478 — #20

478

Figure TF21-3: A full-color thermal-infrared image of a soldier.

industrial and mechanical applications (damaged/failing/ inefficient parts often heat up). This also is used to locate problems in electrical circuits at the board or chip level (Fig. TF21-4). Night vision is important for security, and is highly valued by outdoor enthusiasts as well (IR wildlife

TECHNOLOGY BRIEF 21: SEEING WITHOUT LIGHT cameras can catch pictures of animals when they are most likely to be moving around at night). IR is used for things other than imaging, too, including motion detection and measuring body temperature. Historically, two approaches have been pursued to “see in the dark”: one that relies on measuring self-emitted thermal energy by the scene and another that focuses on intensifying the light reflected by the scene when illuminated by very weak sources, such as the moon or the stars. We will explore each of the two approaches briefly. The visible spectrum extends from the violet (wavelength λ ≈ 0.38 μm) to the red (≈ 0.78 μm). The spectral region next to the visible is the infrared (IR), and it is subdivided into the near-IR (≈ 0.7 to 1.3 μm), mid-IR (1.3 to 3 μm), and thermal-IR (3 to 30 μm). Infrared waves cannot be perceived by humans, because our eyes are not sensitive to EM waves outside of the visible spectrum. In the visible spectrum, we see or image a scene by detecting the light reflected by it, but in the thermal-IR region, we image a scene without an external source of energy, because the scene itself is the source. All material media emit electromagnetic energy all of the time—with hotter objects emitting more than cooler objects. The amount of energy emitted by an object and the shape of its emission spectrum depend on the object’s

Figure TF21-4: Circuit board with superimposed IR image (inset) identifying (in red) high-temperature components or connections. (Credit: Suljo.)

“book” — 2015/5/4 — 7:19 — page 479 — #21

TECHNOLOGY BRIEF 21: SEEING WITHOUT LIGHT

479

1015

Emitted Power Density (W/m3)

1010

6000oC 300oC

105 0oC 100

10−5

10−10 −10 10

10−8

10−6

Ultraviolet and X-ray Optical

10−4

10−2

Infrared

100

102

104

λ (m)

Radio

Wavelength λ (m) Figure TF21-5: Spectra of power density emitted by ideal blackbodies at 0 ◦ C, 300 ◦ C, and 6000 ◦ C.

temperature and its material properties. Most of the emitted energy occurs over a relatively narrow spectral range, as illustrated in Fig. TF21-5, which is centered around a peak value that is highly temperature dependent. For a high-temperature object like the sun (≈ 6000 ◦ C), the peak value occurs at about 0.5 μm (red-orange color), whereas for a terrestrial object, the peak value occurs in the thermal-IR region. Through a combination of lenses and a 2-D array of infrared detectors, the energy emitted by a scene can be focused onto the array, thereby generating an image of the scene. The images sometimes are displayed with a rainbow coloring—with hotter objects displayed in red and cooler objects in blue. In the near- and mid-IR regions, the imaging process is based on reflection—just as in the visible. Interestingly, the sensor chips used in commercial digital cameras are sensitive not only to visible light but to near-IR energy as well. To avoid image blur caused by the IR energy, the camera lens usually is coated with an IR-blocking film that filters out the IR energy but passes visible light with near-perfect transmission. TV remote controls use

near-IR signals to communicate with TV sets, so if an inexpensive digital camera with no IR-blocking coating is used to image an activated TV remote control in the dark, the image will show a bright spot at the tip of the remote control. Some cameras are now making use of this effect to offer IR-based night-vision recording. These cameras emit IR energy from LEDs mounted near the lens, so upon reflection by a nighttime scene, the digital camera is able to record an image “in the dark.”

Image Intensifier A second approach to nighttime imaging is to build sensors with much greater detection sensitivity than the human eye. Such sensors are called image intensifiers. Greater sensitivity means that fewer photons are required in order to detect and register an input signal against the random “noise” in the receiver (or the brain in the case of vision). Some animals can see in the dark (but not in total darkness) because their eye receptors and neural networks require fewer numbers of photons than humans to generate an image under darker conditions. Image

“book” — 2015/5/4 — 7:19 — page 480 — #22

480

TECHNOLOGY BRIEF 21: SEEING WITHOUT LIGHT

Figure TF21-6: Schematic of image intensifier assembly and operation.

intensifiers work by a simple principle (Fig. TF21-6). Incident photons (of which there are relatively few in a dark scene) are focused through lenses and onto a thin plate of gallium arsenide material. This material emits one electron every time a photon hits it. Importantly, these electrons are emitted at the locations where the photons hit the plate, preserving the shape of the light image. These photoelectrons then are accelerated by a high voltage (∼ 5000 V) onto a microchannel plate (MCP). The MCP is a plate that emits 10,000 new electrons every time one electron impacts its surface. In essence, it is an amplifier with a current gain of 10,000. These secondary electrons again are accelerated—this time onto phosphors that glow when impacted with electrons. This works on the same principle as the cathode ray tube. The phosphors are arranged in arrays and form pixels on a display, allowing the image to be seen by the naked eye.

Visible light

UV image

Figure TF21-7: Comparison of visible light and UV images. The latter shows skin damage. (Credit: Milford MD Advanced Dermatology Pocono Medical Care, Inc.)

Ultraviolet Imaging On the other end of the spectrum, UV wavelengths range from 400–200 nm and beyond. UV can also be used to see things that are out of the visible spectrum, particularly skin or soft tissue damage, as shown in the picture of sun-damaged skin in Fig. TF21-7. Dark areas of the skin show where UV is absorbed and not reflected.This can be used for treatment planning, and also to show people the value of skin protection from the sun. UV is also used for numerous astrophysical observations, including the solar flare image shown in Fig. TF21-8. Much information in the universe is outside of the visible spectrum.

Figure TF21-8: Giant solar flare captured in UV light. (Courtesy NASA/SDO and the AIA, EVE, and HMI science team.)

“book” — 2015/5/4 — 7:19 — page 481 — #23

8-5

MAXIMUM POWER TRANSFER

481

In Section 3-6, we established that for a purely resistive circuit, the power transferred from the source circuit to the load is a maximum when RL = RS . The question we now pose is: What are the equivalent conditions for an ac circuit with complex impedances? To answer the question, we start by writing down the expression for IL (the current flowing into the load) namely Vs Vs = . IL = Zs + Z L (Rs + RL ) + j (Xs + XL )

Incorporating Eq. (8.63) in Eq. (8.61) gives (8.64)

The conditions on XL and RL can be combined into

Vs∗ · RL (Rs + RL ) − j (Xs + XL )

1 |Vs |2 RL = . 2 (Rs + RL )2 + (Xs + XL )2

(8.63)

(8.58)

1 1 |IL |2 RL = IL × IL∗ RL 2 2 1 Vs = 2 (Rs + RL ) + j (Xs + XL ) ×

XL = −Xs .

RL = Rs .

From Eq. (8.39a), the average power transferred to (consumed by) the load is Pav =

which when set equal to zero yields

(8.59)

The load parameters RL and XL represent orthogonal dimensions in the complex plane. Hence, the values of RL and XL that maximize Pav can be obtained by performing independent maximization processes: one by setting ∂Pav /∂RL = 0 and another by setting ∂Pav /∂XL = 0. For RL ,

ZL = Z∗s

(maximum power transfer),

(8.65)

where Z∗s = (Rs − j Xs ) is the complex conjugate of Zs . When the condition represented by Eq. (8.65) is true, the load is said to be conjugate matched to the source. According to the result encapsulated by Eq. (8.65):  The average power transferred to (consumed by) an ac load is a maximum when its impedance ZL is equal to Z∗s , which is the complex conjugate of the Th´evenin impedance of the source circuit.  Under the conditions of maximum power transfer represented by Eqs. (8.63) and (8.64), the expression for Pav given by Eq. (8.59) reduces to

1 |Vs |2

2 + (X + X )2 − 2R (R + R )  P (max) = (8.66) . 1 (R + R ) ∂Pav av s L s L L s L 8 RL = |Vs |2 . 2 2 2 ∂RL 2 [(Rs + RL ) + (Xs + XL ) ] (8.60) The right-hand side of Eq. (8.60) is equal to zero if its numerator is equal to zero (because the other alternative, namely setting Example 8-7: Maximum Power the denominator equal to infinity, produces a solution in which ZL and Zs are open circuits corresponding to no power transfer to the load). That is, Determine the maximum amount of power that can be consumed by the load ZL in the circuit of Fig. 8-13. (Rs + RL )2 + (Xs + XL )2 − 2RL (Rs + RL ) = 0, Solution: We start by determining the Th´evenin equivalent of the circuit to the left of terminals (a, b). In Fig. 8-13(b), which simplifies to the load has been removed so as to calculate the open-circuit voltage. Voltage division yields (8.61) Rs2 − RL2 + (Xs + XL )2 = 0. (4 + j 6) × 24 = 17.31 19.44◦ V, Vs = Voc = Similarly, the partial derivative of Pav with respect to XL is 4 + 4 + j6

 where Vs is the Th´evenin voltage of the source circuit to the −2(Xs + XL ) 1 ∂Pav , (8.62) = |Vs |2 RL 2 2 left of terminals (a, b). The Th´evenin impedance of the source ∂XL 2 (Rs + RL ) + (Xs + XL )

“book” — 2015/5/4 — 7:19 — page 482 — #24

482

CHAPTER 8 AC ANALYSIS

−j3 Ω



24

0o

V

+ _

Concept Question 8-10: To achieve maximum transfer of power from a source circuit to a load, how should the impedance of the load be related to that of the source circuit? (See )

a

j6 Ω ZL 4Ω

Concept Question 8-11: Suppose that a certain passive

circuit—containing resistors, capacitors, and inductors— is connected to a square-wave voltage source. What procedure would you use to analyze the voltages and currents in the circuit? (See )

b

(a) −j3 Ω



a

+ 24

0o

V

+ _

8-6

j6 Ω Voc = 17.31 4Ω

19.44o

V

_ Open-circuit voltage

b

(b) −j3 Ω



a

Zs = ZTh = (2.72 − j2.04) Ω b

(c)

This section introduces Multisim power-measurement tools and demonstrates their ability through an interactive simulation of an impedance-matching network. In Section 8-5 we established that the amount of power transferred to a load from a source is at a maximum when the impedance of the load ZLoad is the complex conjugate of the source impedance ZSource . That is, ZLoad = Z∗Source .

j6 Ω 4Ω

Measuring Power with Multisim

' Thevenin impedance

Figure 8-13: Circuit for Example 8-7.

Consider the circuit shown in Fig. 8-14. The circuit is supplied by a realistic source composed of an ideal voltage source Vs in series with a source resistance Rs . The load is a series RL circuit. In the phasor domain: ZSource = Rs

ZLoad = RL + j ωLL .

and

Rs

4(4 + j 6) Zs = 4  (4+j 6)−j 3 = −j 3 = (2.72−j 2.04) . 4 + 4 + j6

a

CM

RM

c RL

ZSource

+

Vs + _ −

~

ZLoad ZLoad + Match

ZL = Z∗s = (2.72 + j 2.04) ,

Source

and the corresponding value of Pav is |Vs |2 (17.31)2 Pav (max) = = = 13.77 W. 8RL 8 × 2.72

(8.68)

For the general case where LL = 0 and Rs = RL , the load would not be matched to the source, and power transfer would not be a maximum. By inserting a matching network in between the source and the load and selecting the values of

circuit, Zs , is obtained by calculating the impedance at terminals (a, b), as shown in Fig. 8-13(c), after deactivating the 24 V voltage source,

For maximum transfer of power to the load, the load impedance should be

(8.67)

b Matching network

d

LL Load

Figure 8-14: Matching network in between the source and the load.

“book” — 2015/5/4 — 7:19 — page 483 — #25

8-6

MEASURING POWER WITH MULTISIM

483

Figure 8-15: Multisim simulation of matching network (CM, RM) in between the source and the load, and wattmeter displays for maximum power transfer.

its components appropriately, we can match the source to the load, thereby realizing the maximum transfer of power from the source to the circuit segment to the right of terminals (a, b), which includes the matching network and the load. If the load has an inductor, the matching network should have a capacitor, and vice versa. For the circuit to the right of terminals (a, b), which includes both the matching network and the load,   1 . (8.69) ZLoad+Match = (RM + RL ) + j ωLL − ωCM For the maximum transfer of power at terminals (a, b) towards the load, it is necessary that ZSource = Z∗Load+Match ,

(8.70)

which can be satisfied by selecting RM and CM as RM = Rs − RL

and

CM =

1 , ω 2 LL

(8.71)

provided Rs ≥ RL . Under these matched conditions, the impedance of the capacitor cancels out the impedance of LL , and the source is matched to the combination of the matching

network and load. This means that the power transferred from the source to this combination is a maximum, but it does not mean that the power transferred to the load alone is a maximum. In fact, if the values of Rs and RL cannot be changed, power transfer to the load is a maximum when RM = 0 and CM = 1/(ω2 LL ). We also should note that the value of CM required to achieve the matching condition is a function of ω. Thus, if the value of CM is selected so as to match the circuit at a given frequency, the circuit will cease to remain matched if ω is changed to a significantly different value. To serve its intended function with significant flexibility, the matching network usually is configured to include a potentiometer and an adjustable capacitor, allowing for manual tuning of RM and CM to satisfy Eq. (8.71) at any specified value of ω (within a certain range). The circuit in Fig. 8-14 can be simulated and analyzed by Multisim, as shown in Fig. 8-15. For variable components, you can choose which keys will shift the component values by double-clicking the component and selecting the desired key letter under Values → Key. Measurement instruments XWM1 and XWM2 are wattmeters configured to measure the average

“book” — 2015/5/4 — 7:19 — page 484 — #26

484

CHAPTER 8 AC ANALYSIS

S

Figure 8-16: Multisim circuit without instruments. power dissipated by a component or circuit: Pav =

1 2

Re[VI∗ ],

where V is the phasor voltage across the component or circuit and I is the phasor current flowing into its positive voltage terminal. In Fig. 8-15, XWM2 measures the current through Rs and the voltage across it, and XWM1 measures the voltage at node 7 (relative to the ground terminal) and the current through the loop at node 7. Thus, XWM2 measures the average power dissipated in Rs , and XWM1 measures the average power delivered by the source to the matching network and load combined. To match the load to the source in the circuit of Fig. 8-15, we should select RM = Rs − RL = 50 − 25 = 25 and CM =

1 1 = = 25.33 μF. ω 2 LL (2π × 103 )2 × 10−3

In Fig. 8-15, RM is a 50 potentiometer set at 50 percent of its maximum value (or 25 ), and CM is a variable 50 μF capacitor, also set at 50 percent of its maximum value (which is very close to the required value of 25.33 μF). The wattmeter displays confirm that the average powers reported by XWM1 and XWM2 are indeed equal. It is important to note that the wattmeter calculates the average power by measuring the voltage and current at a sampling rate specified by the Maximum Time Step (TMAX) in the Interactive Simulation Settings. The default value is

10−5 s, which means that the voltage and current are sampled at a time spacing of 10−5 s. At 1 kHz, the period is 10−3 s. Hence at a time spacing of 10−5 s each cycle gets sampled 100 times, which is quite adequate for generating a reliable measurement of the average power. At higher oscillation frequencies, however, the period is much shorter necessitating that TMAX be selected such that TMAX ≤ 10−2 /f where f is the oscillation frequency in Hz. Thus, at f = 1 MHz, for example, TMAX should be set at 10−8 s. Another method for measuring average power in Multisim is to use the Analysis functions to plot the complex power across any section of a circuit. Figure 8-16 is a Multisim reproduction of the circuit in Fig. 8-15 but with no instruments and fixedvalue components. Note that to perform the AC Analysis Simulation properly, the AC Analysis Magnitude value of the VS source must be changed to 2.5 ∗ sqrt(2) = 3.5355 V. We can plot the magnitude and phase of the complex power S across terminals (3,0) in Fig. 8-16 by performing AC Analysis in Multisim. Under Simulate → Analyses → AC Analysis, set FSTART to 1 Hz and FSTOP to 1 MHz. Make sure to include at least 10 points per decade to produce a good plot. Under Output, enter the following expression: 0.5*(real(I(v1)),-imag(I(v1)))*V(3). Note that this expression is equivalent to S = 21 I∗ V (Eq. (8.29)). (The expression (real(X),-imag(X)) gives us the complex conjugate of any complex number X; we need to do this because Multisim does not have a complex conjugate function). Figure 8-17 shows a plot of the AC Analysis output. As expected, the phase of S goes to 0 at 1 kHz (since it is at this frequency that the inductor and capacitor reactances cancel each other out).

“book” — 2015/5/4 — 7:19 — page 485 — #27

8-6

MEASURING POWER WITH MULTISIM

485

Concept Question 8-12: How is power measured in

Multisim? Why must all four terminals of the wattmeter be used to obtain a power measurement? (See ) Concept Question 8-13: Assuming the values of Vs, Rs,

RL, and LL are fixed, what values of RM and CM lead to maximum transfer of power from the source to RL? (See )

Exercise 8-6: Use Multisim to simulate the circuit in Fig. 8-15. Connect Channel B of the oscilloscope across the voltage source Vs . Vary CM over its full range, noting the phase difference between the two channels of the oscilloscope at CM = 0, CM = 25 μF, and CM = 50 μF.

Figure 8-17: Spectral plots of the magnitude and phase of the complex power S at terminals (3,0) in Fig. 8-16.

Answer: (See

)

Summary Concepts • Even though the average values of the sinusoidal voltage across and current through a load are both zero, the average power consumed by the load is not zero, unless the load is purely reactive (no resistors). • Power is characterized by several attributes, including the complex power S, the average power Pav , and reactive power Q. • The power factor pf is the ratio of the average real power Pav consumed by the load to S (the magnitude of the complex power) which incorporates the reactive 2 + Q2 ]1/2 . power Q through S = [Pav

• An RL load can be compensated by adding a shunt capacitor, causing its pf to increase, and in turn reducing the amount of current that has to be supplied by the electrical power source. • The power transferred from an input source circuit with Th´evenin impedance Zs = Rs + j Xs to a complex load with impedance ZL = RL +j XL is at a maximum when ZL = Z∗s . • Multisim can be used to measure the magnitude and phase of complex power as a function of frequency.

Mathematical and Physical Models Average value

rms value

Xav

1 = T

T x(t) dt 0

Xrms = Xeff

Power factor

   T 1 = x 2 (t) dt T 0

Average power

Pav = Vrms Irms cos(φυ − φi )

Complex power

S=

Reactive power

Q = Vrms Irms sin(φυ − φi ) (VAR)

1 2

VI∗

(W)

(VA)

pf =

Pav = cos(φυ − φi ) S

Power factor lead or lag

Table 8-2

Maximum power transfer

ZL = Z∗s

Maximum power

Pav (max) =

1 |Vs |2 8 RL

“book” — 2015/5/4 — 7:19 — page 486 — #28

486

CHAPTER 8 AC ANALYSIS

Important Terms

Provide definitions or explain the meaning of the following terms:

apparent power average power average value compensated load compensated load circuit complex power consume effective value

impedance matching network in phase instantaneous current instantaneous power instantaneous voltage lagging leading matched load

PROBLEMS

mean periodicity property power factor power factor angle power factor compensation reactive power root root-mean-square (rms) value

square store VAR volt-ampere volt-ampere reactive

8.3 Determine (a) the average and (b) rms values of the periodic current waveform shown in Fig. P8.3.

Section 8-1: Periodic Waveforms *8.1 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.1.

i (A) 4

υ (V)

3 4

2

3

1

2

t (s) 0

1 t (s) 0

1

2

3

4

5

6

7

1

2

3

4

5

6

7

8

Figure P8.3: Waveform for Problem 8.3.

8

Figure P8.1: Waveform for Problem 8.1.

8.2 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.2.

*8.4 Determine (a) the average and (b) rms values of the periodic current waveform shown in Fig. P8.4.

υ (V)

i (A)

4

6

3

4

2

2 t (s)

1

0

1

2

3

4

5

6

7

8

t (s) 0

1

2

3

4

5

6

7

8

Figure P8.4: Waveform for Problem 8.4.

Figure P8.2: Waveform for Problem 8.2.



Answer(s) available in Appendix G.

8.5 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.5.

“book” — 2015/5/4 — 7:19 — page 487 — #29

PROBLEMS

487

υ (V)

i(A) 4

8

2

6 4

0

2

2 5

8

t (s)

10

−2

t (s) 0

6

4

10 15 20 25 30 35 40

−4

Figure P8.5: Waveform for Problem 8.5. Figure P8.8: Waveform for Problem 8.8.

8.6 Determine (a) the average and (b) rms values of the periodic current waveform shown in Fig. P8.6.

8.9 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.9.

i (A) 6 0

υ(V) 4

8

12 16 20 24 28 32 t (s)

4 3

−6 2 1.5 1

−12 Figure P8.6: Waveform for Problem 8.6.

0

0

4

5

6

7

8

t (s)

υ(V)

3t2

(2, 5)

5

6

4

3

3 t (s) 0

3

*8.10 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.10.

υ (V)

9

2

Figure P8.9: Waveform for Problem 8.9.

*8.7 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.7.

12

1

1

2

3

4

5

6

7

8

Figure P8.7: Waveform for Problem 8.7.

8.8 Determine (a) the average and (b) rms values of the periodic current waveform shown in Fig. P8.8.

(6, 5) (4, 3)

(8, 3)

2 1 0

0

2

4

6

8

Figure P8.10: Waveform for Problem 8.10.

t (s)

“book” — 2015/5/4 — 7:19 — page 488 — #30

488

CHAPTER 8 AC ANALYSIS

8.11 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.11.

υ(V) 2

υ(V)

1 1

2

3

4

5

6

0

t (s)

0

2

4

6

8

t (s)

−1

−1

−2 Figure P8.11: Waveform for Problem 8.11.

 υ(t) =

*8.12 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.12.

υ(V)

1 3 4 t V, 1 3 4 (t − 4)

−2 ≤ t ≤ 2 s V, 2 ≤ t ≤ 6 s

Figure P8.13 Waveform for Problem 8.13.

(d) i2 (t) = 4 sin(10t) − 6 sin2 (5t) A.

1 0.86

8.16 Determine the average and rms values of the following periodic waveforms: (a) υ(t) = |12 cos(ωt + θ)| V (b) υ(t) = 4 + 6 cos(2πf t + φ) V

0

2 1 3 4  0≤t ≤1s (1 − e−2t ) V, υ(t) = −2(t−1) −2 (e − e ) V, 1 ≤ t ≤ 2 s

0

t (s)

Figure P8.12 Waveform for Problem 8.12.

8.13 Determine (a) the average and (b) rms values of the periodic voltage waveform shown in Fig. P8.13. 8.14 The current waveform shown in Fig. P8.8 dissipates average power at a rate of 3.2 kW when connected to a resistor. What is the value of the resistor? 8.15 Determine the average and rms values of the following periodic waveforms: (a) υ1 (t) = 4 cos(60t − 30◦ ) V. (b) i1 (t) = 2.5 A. *(c) υ2 (t) = 12 − sin(2t + 45◦ ) V.

(c) υ(t) = 2 cos ωt − 4 sin(ωt + 30◦ ) V (d) υ(t) = 9 cos ωt sin(ωt + 30◦ ) V Section 8-2 and 8-3: Average and Complex Power 8.17 Determine the complex power, apparent power, average power absorbed, reactive power, and power factor (including whether it is leading or lagging) for a load circuit whose voltage and current at its input terminals are given by: (a) υ(t) = 100 cos(377t − 30◦ ) V, i(t) = 2.5 cos(377t − 60◦ ) A. (b) υ(t) = 25 cos(2π × 103 t + 40◦ ) V, i(t) = 0.2 cos(2π × 103 t − 10◦ ) A. *(c) Vrms = 110 60◦ V, Irms = 3 45◦ A. (d) Vrms = 440 0◦ V, Irms = 0.5 75◦ A. (e) Vrms = 12 60◦ V, Irms = 2 −30◦ A. 8.18 Determine the complex power, apparent power, average power absorbed, reactive power, and power factor (including whether it is leading or lagging) for a load circuit whose voltage and current at its input terminals are given by:

“book” — 2015/5/4 — 7:19 — page 489 — #31

PROBLEMS

489

(a) υ(t) = 110 cos(60t − 60◦ ) V, i(t) = 4 cos(60t − 25◦ ) A.

8.22 In the phasor-domain shown in Fig. P8.22, V = 120 0◦ V, I = 0.3 30◦ A, ω = 1000 rad/s, R1 = 200 , R2 = 200 , R3 = 1.2 k , L = 0.2 H, and C = 10 μF. Determine the complex power for each passive element, and verify that conservation of energy is satisfied.

(b) υ(t) = 12 cos(1000t + 30◦ ) V, i(t) = 0.4 cos(1000t − 15◦ ) A. (c) υ(t) = −10 cos(3000t + 10◦ ) V, i(t) = 0.5 sin(3000t − 5◦ ) A. (d) Vrms = 240 0◦ V, Irms = 0.8 50◦ A.

R2

R1

(e) Vrms = 6 20◦ V, Irms = 255 15◦ mA. (f) Vrms = 18 40◦ V, Irms = 1 −50◦ A.

V

+ _

C

L

R3

I

8.19 Determine the impedance of a load characterized by the following attributes: (a) S = 1.2 30◦ kVA, Vrms = 40 0◦ V,

Figure P8.22: Circuit for Problem 8.22.

(b) |S| = 80 VA, Q = 26 VAR, Irms = 4 45◦ A. (c) Vrms = 25 −15◦ V, Irms = 0.5 35◦ A. 8.20 In the circuit shown in Fig. P8.20, υ(t) = 40 cos(105 t) V, R1 = 100 , R2 = 500 , C = 0.1 μF, and L = 0.5 mH. Determine the complex power for each passive element, and verify that conservation of energy is satisfied.

*8.23 In the circuit of Fig. P8.23, υs (t) = 60 cos 4000t V, R1 = 200 , R2 = 100 , and C = 2.5 μF. Determine the average power absorbed by each passive element and the average power supplied by the source.

R1

R1 υs(t) υ(t)

+ _

R2

C

R2

+ _

C

Figure P8.23: Circuit for Problem 8.23.

L

Figure P8.20: Circuit for Problem 8.20.

*8.21 In the circuit shown in Fig. P8.21, υ(t) = 12 cos(2000t) V, R = 20 , and C = 4.7 μF. Determine the complex power of the source. What is the power factor of the voltage source.

R

8.24 In the circuit of Fig. P8.24, is (t) = 0.2 sin 105 t A, R = 20 , L = 0.1 mH, and C = 2 μF. Show that the sum of the complex powers for the three passive elements is equal to the complex power of the source.

+ is(t) _

R

L

C

C Figure P8.24: Circuit for Problem 8.24.

υ(t)

+ _

Figure P8.21: Circuit for Problem 8.21.

8.25 In the phasor-domain circuit of Fig. P8.25, Vs = 20 V, Is = 0.3 30◦ A, R1 = R2 = 100 , ZL = j 50 , and ZC = −j 50 . Determine the complex power for each of the four passive elements and for each of the two sources. Verify that conservation of energy is satisfied.

“book” — 2015/5/4 — 7:19 — page 490 — #32

490

CHAPTER 8 AC ANALYSIS

R2

ZC

R1 + Vs _

R2

ZL

Is

L1

a R3

R1

I

C

Load L2

Figure P8.25: Circuit for Problem 8.25.

b Figure P8.28: Circuit for Problem 8.28.

*8.26 Determine the average power dissipated in the load resistor RL of the circuit in Fig. P8.26, given that Vs = 100 V, R1 = 1 k , R2 = 0.5 k , RL = 2 k , ZL = j 0.8 k , and ZC = −j 4 k .

R1

R2

+ Vs _

*8.29 In the phasor-domain circuit shown in Fig. P8.29, V = 100 0◦ V, R1 = 1 k , R2 = 0.6 k , RL = 3 k , ZL = j 0.8 k , and ZC = −j 0.5 k . Determine the average power dissipated in RL .

ZL

ZL

R1

a RL

ZC b

V

+ _

a

R2

RL

ZC b

Figure P8.26: Circuit for Problem 8.26.

Figure P8.29: Circuit for Problem 8.29.

8.27 Determine S for the RL load in the circuit of Fig. P8.27, given that Is = 4 0◦ A, R1 = 10 , R2 = 5 , ZC = −j 20 , R = 10 , and ZL = j 20 .

R2

8.30 In the circuit shown in Fig. P8.30, i(t) = 3 cos(1000t) A, R1 = 2 k , R2 = 560 , RL = 2 k , and L = 0.4 H. Determine the average power dissipated in RL .

4ix

a R2

R ZC

R1

Is ZL b Load

a

i(t)

ix

R1

RL

L b

Figure P8.27: Circuit for Problem 8.27.

Figure P8.30: Circuit for Problem 8.30.

8.28 In the phasor-domain circuit shown in Fig. P8.28, I = 2 0◦ A, R1 = 20 , R2 = 1 , R3 = 5 , ZL1 = j 5 , ZL2 = j 25 , and ZC = −20 . Determine S of the load.

*8.31 In the phasor-domain circuit shown in Fig. P8.31, V = 15 45◦ V, R1 = 5 , R2 = 2 , ZC = −j 1 , ZL = j 2 , and Zload = 6 + j 4 . Determine the complex power of the load.

“book” — 2015/5/4 — 7:19 — page 491 — #33

PROBLEMS

491

ZL

R1

V

I

+ _

ZC

R2

a ZC

R2 + _

R1

_ +

Zload 2I + _

Vin

b

ZL RL

Figure P8.31: Circuit for Problem 8.31.

8.32 In the op-amp circuit shown in Fig. P8.32, υin (t) = 12 cos(1000t) V, R = 10 k , RL = 5 k , and C = 1 μF. Determine the complex power for each of the passive elements in the circuit. Is conservation of energy satisfied?

Figure P8.33: Op-amp circuit for Problem 8.33.

R1

_ +

υin(t)

C

υin(t)

+ _

_ +

+

RL

C2

C1 R

υout(t)

υout(t) Figure P8.34: Op-amp circuit for Problem 8.34.

RL

_ Figure P8.32: Op-amp circuit for Problem 8.32.

+ _

j1.6 Ω

0.5 Ω

I

0.3 Ω

a

0.2 Ω 2

30 V

RL = 0.5 Ω 0.25I b

*8.33 In the phasor-domain op-amp circuit shown in Fig. P8.33, Vin = 2 0◦ V, R1 = 200 , R2 = 2.4 k , RL = 10 k , ZC = −j 500 , and ZL = j 1 k . Determine the average power delivered to RL .

8.35 Determine the power dissipated in RL of the circuit in Fig. P8.35. *8.36 Determine the power dissipated in RL of the circuit in Fig. P8.36.

2IC

j3 Ω IC

3Ω + _ 8

_

45

−j1 Ω

a

+

8.34 In the phasor-domain op-amp circuit shown in Fig. P8.34, υ(t) = 8 cos(200t) V, R1 = 5 k , RL = 2 k , C1 = 1 μF, and C2 = 4.7 μF. Determine the average power delivered to RL .

Figure P8.35: Circuit for Problem 8.35.



RL = 4 Ω

V

Figure P8.36: Circuit for Problem 8.36.

b

“book” — 2015/5/4 — 7:19 — page 492 — #34

492

CHAPTER 8 AC ANALYSIS

8.37

In the op-amp circuit of Fig. P8.37,

8.40 The apparent power entering a certain load Z is 250 VA at a power factor of 0.8 leading. If the rms phasor voltage of the source is 125 V at 1 MHz:

υin (t) = V0 cos ωt V, with V0 = 10 V, ωRC = 1, and RL = 10 k . Determine the power delivered to RL .

__

+ υin(t) _

+

+

υout

RL

Figure P8.37: Op-amp circuit for Problem 8.37.

8.38 Determine the amount of power delivered to RL in the circuit of Fig. P8.38, given that υin (t) = 0.5 cos 2000t V, R1 = 1 k , R2 = 10 k , C = 0.1 μF, RL = 1 k , and L = 0.2 H.

R2

8.41 Voltage source Vs in the circuit of Fig. P8.41 supplies power to three load circuits with impedances Z1 , Z2 , and Z3 . The following partial power information was deduced from measurements performed on the three load circuits: Load Z1 :

80 W at pf = 0.8 lagging

Load Z2 :

60 VA at pf = 0.7 leading

Load Z3 :

40 VA at pf = 0.6 leading

(a) the rms value of Vs by applying the law of conservation of energy

_ +

IL

+ υin(t) _

υout

(b) Z1 , Z2 , and Z3 .

RL

100 Ω + Vs _

L

Figure P8.38: Op-amp circuit for Problem 8.38.

10 kΩ

_

υout

+ 0.1 μF

RL = 5 kΩ

Figure P8.39: Circuit for Problem 8.39.

Irms

Z1

Z2

Z3

Figure P8.41: Circuit for Problem 8.41.

*8.39 Given that υs (t) = 2 cos 103 t V in the circuit of Fig. P8.39, determine the power delivered to RL .

υs(t)

(d) The equivalent impedance of the load circuit should be of the form Z = R + j ωL or Z = R − j/ωC. Determine the value of L or C, whichever is applicable.

If Irms = 0.4 37◦ A, determine:

C R1

(b) Determine S into the load (c) Determine Z

R

C

(a) Determine Irms going into the load

8.42 The apparent power entering a certain load Z is 120 VA at a power factor of 0.866 lagging. If the rms phasor voltage is 240 V at 60 Hz: (a) Determine S into the load (b) Determine Irms going into the load (c) Determine Z (d) The equivalent impedance of the load circuit should be of the form Z = R +j ωL or Z = R +1/j ωC. Determine the value of L or C, whichever is applicable.

“book” — 2015/5/4 — 7:19 — page 493 — #35

PROBLEMS

493

*8.43 In the circuit in Fig. P8.43, voltage source Vs supplies power to three load circuits with impedances Z1 , Z2 , and Z3 . The following information was deduced from measurements performed on the three load circuits: Load Z1 :

100 VA at pf = 0.6 lagging

Load Z2 :

70 VA at pf = 0.75 leading

Load Z3 :

45 W at pf = 0.95 lagging

Section 8-4: Power Factor *8.45 The RL load in Fig. P8.45 is compensated by adding the shunt capacitance C so that the power factor of the combined (compensated) circuit is exactly unity. How is C related to R, L, and ω in that case?

R C

If Vs = 100 0◦ V, determine the equivalent impedance. Is it inductive or capacitive? Are Z1 , Z2 , and Z3 inductive or capacitive?

L

Figure P8.45: Circuit for Problem 8.45.

Vs

+ _

Z1

Z2

Z3

Figure P8.43: Circuit for Problem 8.43.

8.44 In the circuit shown in Fig. P8.44, voltage source Vs supplies power to three load circuits with impedances Z1 , Z2 , and Z3 . The following information was deduced from measurements performed on the three load circuits. Load Z1 :

60 VA at pf = 0.866 lagging

Load Z2 :

80 W at pf = 0.750 leading

Load Z3 :

100 VAR at pf = 0.600 leading

If Irms = 0.5 45◦ and R = 100 , determine: (a) the rms value of Vs , by applying the law of conservation of energy

8.46 The generator circuit shown in Fig. P8.46 (see page 493) is connected to a distant load via a long coaxial transmission line. The overall circuit can be modeled as in Fig. P8.46(b), in which the transmission line is represented by an equivalent impedance Zline = (5 + j 2) . (a) Determine the power factor of voltage source Vs . (b) Specify the capacitance of a shunt capacitor C that would raise the power factor of the source to unity when connected between terminals (a, b). The source frequency is 1.5 kHz. 8.47 Source Vs in the circuit of Fig. P8.47 is connected to two industrial loads, with equivalent impedances Z1 and Z2 , via two identical transmission lines, each characterized by an equivalent impedance Zline = (0.5 + j 0.3) . If Z1 = (8 + j 12) and Z2 = (6 + j 3) : (a) Determine the power factors for Z1 , Z2 , and source Vs . (b) Specify the capacitance of a shunt capacitor C that would raise the power factor of the source to 0.95 when connected between terminals (a, b). The source frequency is 12 kHz.

(b) Z1 , Z2 , and Z3 .

Z2

R

Vs

+ _



Irms Z1

Z3

Transmission line a 0.5 Ω j0.3 Ω

+ Vs _

C

Transmission line 0.5 Ω j0.3 Ω Z1

b Figure P8.44: Circuit for Problem 8.44.

Figure P8.47: Circuit for Problem 8.47.

Z2

“book” — 2015/5/4 — 7:19 — page 494 — #36

494

CHAPTER 8 AC ANALYSIS

a

10 Ω

Vs

c

+ _

ZL = (50 + j40) Ω Transmission line

b

d

(a) Transmission-line circuit

a

10 Ω

j2 Ω



c

Transmission line Vs

+ _

ZL = (50 + j40) Ω

C

(b) Equivalent circuit

b

d Figure P8.46: Circuit for Problem 8.46.

*8.48 Use the power information given for the circuit in Fig. P8.48 to determine: (a) Z1 and Z2 (b) the rms value of Vs .

0.6 Ω + _ Vs

1.2 Ω

j0.4 Ω

+ Z1

Z2

Vrms = 440

_

(a) Determine the power factor of the load, the power factor of the transmission line, and the power factor of the voltage source. (b) Specify the capacitance of a shunt capacitor C that would raise the power factor of the source to unity when connected between terminals (a, b). The source frequency is 60 Hz.

0 V

Rs Load Z1 : 24 kW @ pf = 0.66 leading Load Z2 : 18 kW @ pf = 0.82 lagging Vs

a

+ _

Transmission line

C

Figure P8.48: Circuit for Problem 8.48.

8.49 In the circuit shown in Fig. P8.49, a generator is connected to a load via a transmission line. Given that Rs = 10 , Zline = (4 + j 2) , and Zload = (40 + j 30) :

b Figure P8.49: Circuit for Problem 8.49.

Load

“book” — 2015/5/4 — 7:19 — page 495 — #37

PROBLEMS

495

8.50 In the phasor-domain circuit shown in Fig. P8.50, R1 = 1.2 k , R2 = 5 k , L1 = 0.8 H, L1 = 0.6 H, C = 2 μF, and ω = 1000 rad/s.

8.53 For the circuit in Fig. P8.53, choose the load impedance ZL so that the power dissipated in it is a maximum. How much power will that be?

*(a) Determine the power factor of the voltage source V. (b) Specify the capacitance of a shunt capacitor Cshunt that would raise the power factor at terminals (a, b) to 0.9 when connected between terminals (a, b).

+ _

V

C

L1

a

Cshunt

2I

j2 Ω

_

L2

R2

8

j6 Ω

I

1Ω 0

+

R1



ZL

−j1 Ω

+ V _

Figure P8.53: Circuit for Problem 8.53.

b Figure P8.50: Circuit for Problem 8.50.

8.54 For the circuit in Fig. P8.54, choose the load impedance ZL so that the power dissipated in it is a maximum. How much power will that be?

Section 8-5: Maximum Power Transfer 8.51 For the circuit in Fig. P8.51, choose the load impedance ZL so that the power dissipated in it is a maximum. How much power will that be?

−j2 kΩ

4 kΩ 2

0

mA

−j4 Ω

4Ω 20

0

+ V _

j2 Ω

ZL −j6 kΩ

8 kΩ

ZL

Figure P8.54: Circuit for Problem 8.54.

Figure P8.51: Circuit for Problem 8.51.

*8.52 For the circuit in Fig. P8.52, choose the load impedance ZL so that the power dissipated in it is a maximum. How much power will that be?

*8.55 For the circuit in Fig. P8.55, choose the load impedance ZL so that the power dissipated in it is a maximum. How much power will that be?

−j2 Ω j4 kΩ j2 Ω

1Ω 6

0

V

+ _

3 kΩ ZL

Figure P8.52: Circuit for Problem 8.52.



15

0

+ V _

2000Ix +_

Ix 6 kΩ

Figure P8.55: Circuit for Problem 8.55.

ZL

“book” — 2015/5/4 — 7:19 — page 496 — #38

496

CHAPTER 8 AC ANALYSIS

8.56 In the phasor-domain shown in Fig. P8.56, V = 20 0◦ V, R1 = 10 , R2 = 5 , ZC = −j 5 , and ZL = j 3 . Choose the load impedance Zload so that the power dissipated in it is a maximum. How much power will that be?

2I _

+

I

ZL

V

V

R3

ZL1

R1 R1

R2

ZL 2 Zload

ZC

+ _

R2

+ _

ZC

Figure P8.58: Circuit for Problem 8.58.

Zload 8.59 In the phasor-domain circuit shown in Fig. P8.59, V = 12 0◦ V, R1 = 2 k , R2 = 4 k , ZL1 = j 2 k , and ZL2 = j 5 k . Choose the load impedance Zload so that the power dissipated in it is a maximum. How much power will that be?

Figure P8.56: Circuit for Problem 8.56.

8.57 In the phasor-domain circuit shown in Fig. P8.57, V = 30 60◦ V, R1 = 5 , R2 = 20 , R3 = 10 , ZC = −j 4 , and ZL = j 6 . Choose the load impedance Zload so that the power dissipated in it is a maximum. How much power will that be?

ZL1

R1 V _

+

Zload

R2 R2

Figure P8.59: Circuit for Problem 8.59.

ZC

R1

V

+ _

ZL2

ZL Zload R3

*8.60 In the phasor-domain circuit shown in Fig. P8.60, I = 2.5 0◦ mA, R1 = 10 k , R2 = 2.4 k , R3 = 10 k , and ZC = −j 4 k . Choose the load impedance Zload so that the power dissipated in it is a maximum. How much power will that be?

0.001 VC Figure P8.57: Circuit for Problem 8.57.

R2

I *8.58 In the phasor-domain circuit shown in Fig. P8.58, V = 6 0◦ V, R1 = 1 , R2 = 2 , R3 = 5 , ZL1 = j 2 , ZL2 = j 5 , and ZC = −j 6 . Choose the load impedance Zload so that the power dissipated in it is a maximum. How much power will that be?

R1

ZC

+ _VC

R3

Figure P8.60: Circuit for Problem 8.60.

Zload

“book” — 2015/5/4 — 7:19 — page 497 — #39

PROBLEMS

497

50 Ω

1 μH

+ υs(t) _

ZL = [50 + jω(1 μH)] Ω

1 nF

Figure P8.61: Circuit for Problem 8.61.

1 μH

50 Ω + υs(t) _

Section 8-6: Multisim 8.61 Model the circuit in Fig. P8.61 in Multisim and plot the complex power through the load ZL as a function of frequency from 1 kHz to 1 GHz. Assume υs (t) has an amplitude of 1 V.

1 μH

Zin

1 μF

Figure P8.62: Circuit for Problems 8.62 and 8.65.

8.62 Model the circuit in Fig. P8.62 in Multisim and find the frequency at which the input impedance of the load circuit Zin is purely real. Assume υs (t) has an amplitude of 1 V.

8.64 Model the circuit in Fig. P8.64 and use the wattmeter to determine the average power consumed by the load ZL . Also, perform an AC Analysis from 100 kHz to 1 GHz and show that the average power value given by the AC Analysis at 1 MHz matches the value provided by the wattmeter.

8.63 Model the circuit in Fig. P8.63 in Multisim and find the frequency at which the input impedance of the load circuit Zin is purely real.

R 12.5 Ω

C2

L

L + υs(t) _

Zin

R

C3 C1

L

R

R = 50 Ω C1 = 1 μF C2 = 2 μF C3 = 3 μF L = 1 mH

Figure P8.63: Circuit for Problem 8.63.

25 Ω 1 V 0 1 MHz

+ _

1 nF

50 Ω

12.5 Ω 1 μH

 ZL = 12.5 −

Figure P8.64: Circuit for Problem 8.64.

j ω × 1 nF



“book” — 2015/5/4 — 7:19 — page 498 — #40

498

8.65 Plot the power factor and phase angle φz across the load Zin in Fig. P8.62 using AC Analysis in Multisim from 1 kHz to 1 MHz. (See Multisim Demo 8.3 in the Tutorial for help on how to do this.)

CHAPTER 8 AC ANALYSIS

υ1(t) Vm

Potpourri Questions 8.66 What is the wavelength range of visible light? The corresponding frequency range? 8.67 What is the frequency range assigned to Bluetooth communication? 8.68 What does “thermal” imaging refer to? How is ultraviolet light used in skin treatment?

T/4

T/2

3T/4

T

3T/4

T

t

−Vm υ2(t) Vm

T/4

T/2

t

Integrative Problems: Analytical / Multisim / myDAQ

−Vm To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m8.1 Periodic Waveforms: Plot the half-wave rectifier output and full-wave rectifier output for each of the three standard waveforms shown in Fig. m8.1. Then: (a) Determine the general expressions for the (1) average value and (2) rms value for each of the six rectified waveforms.

υ3(t) Vm

T/4

T/2

3T/4

T

−Vm Figure m8.1 Circuit for Problem m8.1.

(b) Evaluate your expressions for average and rms values for Vm = 10 V and T = 10 ms. m8.2 Average Power: The circuit shown in Fig. m8.2 operates in sinusoidal steady state at 1500 Hz. The voltage source amplitude is 3 V. Find the average power delivered by the source. Use these component values: R = 100 , C = 1.0 μF, and L = 3.3 mH. m8.3 Complex Power: The circuit shown in Fig. m8.3 operates in sinusoidal steady state at 1000 Hz. The voltage source amplitude is 2.5 V. Component values are: R = 100 , C = 1.0 μF, and L = 3.3 mH.

R

υsrc

~+_

L C

(a) Find the complex power in rectangular format for each of the four circuit elements: Ssrc , SR , SL , and SC . (b) Demonstrate conservation of complex power with these four values.

Figure m8.2 Circuit for Problem m8.2.

t

“book” — 2015/5/4 — 7:19 — page 499 — #41

PROBLEMS

499

R

υsrc

+ _

~

iL

L

υsrc

+ _

~

+

R1

R2

L1

L2

R3

υL

_

C Figure m8.4 Circuit for Problem m8.4. Figure m8.3 Circuit for Problem m8.3.

m8.4 The Power Factor: The circuit shown in Fig. m8.4 is a “scale model” of two industrial electric motors and a heating unit connected to a manufacturing plant power distribution network. The resistor/inductor combinations, R1 -L1 and R2 -L2 , model the winding resistance and magnetic fields of the motors. Resistor R3 models the heater coils. C represents the power factor compensation equipment—essentially a capacitor bank with high power capacity. (a) Determine the power factor of the uncompensated load, and draw its power triangle to scale. (b) Determine the value of the compensation capacitor C required to improve the load power factor to 0.90 lagging.

(c) Available power factor compensation capacitors include 0.1 μF, 1.0 μF, and 10 μF; the cost of compensation equipment increases with capacitance. Choose the least expensive compensation capacitor closest to C and then determine the power factor and power triangle (also drawn to scale) of the compensated load. Component values are: R1 = 10 , R2 = 100 , R3 = 100 , L1 = 3.3 mH, L2 = 33 mH, and Vsrc = 1 V at 2500 Hz (actual industrial motors operate at hundreds of volts and 50 Hz to 60 Hz).

“book” — 2015/5/4 — 7:20 — page 500 — #1

9

CHAPTER

Frequency Response of Circuits and Filters Contents 9-1 9-2 TB22 9-3 9-4 9-5 TB23 9-6 9-7 TB24 9-8 9-9

Overview, 501 The Transfer Function, 501 Scaling, 507 Noise-Cancellation Headphones, 509 Bode Plots, 512 Passive Filters, 522 Filter Order, 530 Spectral and Spatial Filtering, 533 Active Filters, 536 Cascaded Active Filters, 538 Electrical Engineering and the Audiophile, 544 Application Note: Modulation and the Superheterodyne Receiver, 547 Spectral Response with Multisim, 550 Summary, 555 Problems, 556

dB

60

57 M1 [dB] −20 dB/decade (1st order)

50 40

M2 [dB] −40 dB/decade (2nd order)

30 20

M3 [dB] −60 dB/decade (3rd order)

10

Objectives Learn to: 

Derive the transfer function of an ac circuit.

 

Generate magnitude and phase spectral plots. Design first-order lowpass, highpass, bandpass, and bandreject filters.



Generate Bode plots for any transfer function.



Design active filters.



Apply Multisim to generate spectral responses for passive and active circuits.

0 104

ωc3

105 ωc2

ωc1

106

107

ω (rad/s)

Frequency filters are used to suppress noise, remove interfering signals, and to channel multifrequency signals along their intended paths.

“book” — 2015/5/4 — 7:20 — page 501 — #2

9-1 THE TRANSFER FUNCTION

Overview In Chapter 7 we learned how to analyze an ac circuit excited by an individual source at a particular frequency f . Often, the input signal is a superposition of many sinusoidal signals at different frequencies. A good analogue is sunlight incident on an eye’s pupil. The light consists of many ac-like signals at many frequencies, extending from below the frequency of violet to beyond that of red. The pupil is like a receiver that detects the incident light, but it is also a filter because it detects only a portion of the spectrum incident upon it, while rejecting other bands such as the ultraviolet and infrared. Additional filtering can be effected through the use of tinted sunglasses, or similar optical filters. In digital photography, filtering can be performed in software during post-processing (using PhotoshopTM , for example) to enhance certain colors of interest over others. This chapter is about how to design RLC circuits that can filter in (pass through) the range of frequencies of interest (to a certain application, such as in a communication, imaging, or sensing system) and filter out (reject) the range of frequencies of signals that are either problematic or not of interest. To avoid interference, every radio and TV transmission station is assigned a unique transmission frequency different from those assigned to other radio and TV stations in the area. At the receiver end, even though the antenna will intercept the signals transmitted by all sources within a certain distance, the receiver is able to select from among them the specific channel of interest, while rejecting all others. The selection process is based on the oscillation frequency of the desired signal, and it is realized by passing the intercepted signals through a narrow bandpass filter whose center frequency is aligned with the frequency of the desired channel. The bandpass filter is one of many different types of frequency-selective circuits employed in analog and digital communication networks to manage the traffic of signals between multiple sources and multiple recipients. The behavior of an ac circuit as a function of the angular frequency ω is called its frequency response. Building on the phasor-domain analysis tools we acquired in the preceding two chapters, we now are ready to develop and adopt a standard set of metrics and design methodologies for characterizing the frequency response of any resonant circuit and to apply them to various types of active and passive circuits.

9-1 The Transfer Function The passive linear circuit represented by the block diagram in Fig. 9-1 has an input phasor voltage Vin applied at input terminals (a, b), causing an associated input phasor current Iin to flow into the circuit. In general, a corresponding set of

501 phasors, Vout and Iout , exist at output terminals (c, d). The voltage gain of the circuit is defined as H(ω) =

Voltage gain:

Vout (ω) , Vin (ω)

(9.1)

where all quantities are written explicitly as functions of the angular frequency ω simply to emphasize the notion that ω will play a central role in our forthcoming discussions. If the circuit contains capacitors and inductors, Vout likely will be a function of ω, and in the general case Vin may vary with ω also. The phasor H(ω) is called the voltage transfer function of the circuit and carries a connotation broader than just another name for voltage gain. In fact, H(ω) can be defined to convey the relationship between any input excitation and any output response. For example, we may define other transfer functions for the circuit in Fig. 9-1, such as: Iout (ω) , Iin (ω) Vout (ω) HZ (ω) = , Iin (ω) HI (ω) =

Current gain: Transfer impedance:

(9.2a) (9.2b)

and Transfer admittance:

HY (ω) =

Iout (ω) . Vin (ω)

(9.2c)

In any case, because H(ω) always is defined as the ratio of an output quantity to an input quantity, we may think of it as equal to the output generated by the circuit in response to a unity input (1 0◦ ). As a complex quantity, the transfer function H(ω) has a magnitude—to which we assign the symbol M(ω)—and an associated phase angle φ(ω), H(ω) = M(ω) ej φ(ω) ,

(9.3)

where by definition, M(ω) = |H(ω)| and

a

φ(ω) = tan−1



Iin

+ _

Vin

 Im[H(ω)] . (9.4) Re[H(ω)]

Iout H(ω) Linear circuit

b

c

+ V _ out d

Figure 9-1: The voltage-gain transfer function is H(ω) = Vout (ω)/Vin (ω)

.

“book” — 2015/5/4 — 7:20 — page 502 — #3

502

CHAPTER 9

M

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

M

Lowpass

M0

Highpass

M0 Idealized

Idealized

1 M0 √2

1 M0 √2 Slope = Sg Actual

Actual Passband 0

Passband ω

ωc

0

0

(a) Lowpass filter M

ω

ωc

0

(b) Highpass filter M

Bandpass

M0

Bandreject

M0 Idealized Idealized

1 M 0 √2

1 M 0 √2 Actual Actual

Stopband Passband 0

0

ωc1

ω0

Passband

Passband

ωc2

ω

(c) Bandpass filter

0

0

ωc1

ω0

ωc2

ω

(d) Bandreject filter

Figure 9-2: Typical magnitude spectral responses for the four types of filters.

9-1.1 Terminology The voltage transfer functions most commonly encountered in electronic circuits are those belonging to lowpass, highpass, bandpass, and bandreject filters. To visualize the frequency response of a transfer function, we usually generate plots of

its magnitude and phase angle as a function of frequency from ω = 0 (dc) to ω = ∞. Figure 9-2 displays typical magnitude responses for the four aforementioned types of filters. Each of the four filters is characterized by at least one passband and one stopband.

“book” — 2015/5/4 — 7:20 — page 503 — #4

9-1 THE TRANSFER FUNCTION

503

 The lowpass filter allows low-frequency signals to pass through (essentially unimpeded) but blocks the transmission of high-frequency signals. 

Since M(ω) is a voltage transfer function, M 2 (ω) is the transfer function for power. The condition described by Eq. (9.5) is equivalent to M 2 (ωc ) =

The qualifiers low and high are relative to the corner frequency ωc (Fig. 9-2(a)), which we shall define shortly.  The highpass filter exhibits the opposite behavior, blocking low-frequency signals while allowing high frequencies to go through.  The bandpass filter (Fig. 9-2(c)) is transparent to signals whose frequencies are within a certain range centered at ω0 , but cuts off both very high and very low frequencies. The response of the bandreject filter provides the opposite function to that of the bandpass filter; it is transparent to low- and high-frequency signals and opaque to intermediate-frequency signals. We often use the term “frequency” for both the angular frequency ω and the oscillation frequency f = ω/2π . Because the impedances of inductors and capacitors are given by j ωL and 1/j ωC, it is easier to analyze a circuit and plot its response as a function of ω, but if the circuit performance is specified in Hz, ω should be replaced with 2πf everywhere.

M02 2

or

P (ωc ) =

P0 . 2

(9.6)

Hence, ωc also is called the half-power frequency. The spectra of the lowpass and highpass filters shown in Fig. 9-2(a) and (b) have only one half-power frequency each, but the bandpass and bandreject responses have two half-power frequencies each, ωc1 and ωc2 . Even though the actual frequency response of a filter is a gently varying curve, it usually is approximated to that of an equivalent idealized response, as illustrated in Fig. 9-2. The idealized version for the lowpass filter has a rectangle-like envelope with a sudden transition at ω = ωc . Accordingly, ωc also is referred to as the cutoff frequency of the filter. This term also applies to the other three types of filters. Bandwidth B For lowpass and bandpass filters, the bandwidth B is defined as the range of ω corresponding to the filter’s idealized passband (Fig. 9-2):

Gain factor M0

B = ωc

for lowpass filter,

(9.7a)

All four spectral plots shown in Fig. 9-2 exhibit smooth patterns as a function of ω, and each has a peak value M0 in its passband. If M0 occurs at dc, as in the case of the lowpass filter, it is called the dc gain; if it occurs at ω = ∞, it is called the highfrequency gain; and for the bandpass filter, it is called simply the gain factor. In some cases, the transfer function of a lowpass or highpass filter may exhibit a resonance behavior that manifests itself in the form of a peaking pattern in the neighborhood of the resonant frequency of the circuit, ω0 , as illustrated in Fig. 9-3. Obviously, the peak value at ω = ω0 exceeds M0 , but we will continue to refer to M0 as the dc gain of M(ω) because M0 is defined as the reference level in the passband of the transfer function, whereas the behavior of M(ω) in the neighborhood of ω0 is specific to that neighborhood.

B = ωc2 − ωc1

for bandpass filter.

(9.7b)

M 2M0

strong resonance M0 dc gain weak resonance

Corner frequency ωc The corner frequency ωc is√ defined as the angular frequency at which M(ω) is equal to 1/ 2 of the reference value M0 , M0 M(ωc ) = √ = 0.707M0 . 2

(9.5)

0 0

ω0

2ω0

3ω0

ω

Figure 9-3: Resonant peak in the spectral response of a lowpass filter circuit.

“book” — 2015/5/4 — 7:20 — page 504 — #5

504

CHAPTER 9

Resonant frequency ω0  Resonance is a condition that occurs when the input impedance or input admittance of a circuit containing reactive elements is purely real, and the angular frequency at which it occurs is called the resonant frequency ω0 .  Often (but not always) the transfer function H(ω) also is purely real at ω = ω0 , and its magnitude is at its maximum or minimum value. Let us consider the two circuits shown in Fig. 9-4. The input impedance of the RL circuit is simply Zin1 = R + j ωL.

(RL circuit).

R +

~+−_

+

Zin1

L

VL

_

(a) First-order RL filter C

Vs

+ + _ −

~



Zin2

1 = R + j ωL − ωC

 .

(9.10)

At resonance (ω = ω0 ), the imaginary part of Zin2 is equal to zero. Thus, ω0 L −

1 = 0, ω0 C

or ω0 = √

(9.9)

When ω0 = 0 (dc) or ∞, the resonance is regarded as a trivial resonance because it occurs at the extreme ends of the spectrum. This usually happens when the circuit has either an inductor or a capacitor (but not both simultaneously). A circuit

Vs

that exhibits only a trivial resonance, such as the RL circuit in Fig. 9-4(a), is not considered a resonator. If the circuit contains at least one capacitor and at least one inductor, resonance can occur at intermediate values of ω. A case in point is the series RLC circuit shown in Fig. 9-4(b). Its input impedance is

(9.8)

Resonance corresponds to when the imaginary part of Zin1 is zero, which occurs at ω = 0. Hence, the resonant frequency of the RL circuit is ω0 = 0

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

1 LC

(RLC circuit).

(9.11)

So long as neither L nor C is zero or ∞, the transfer function H(ω) = VR /Vs will exhibit a two-sided spectrum with a peak at ω0 —similar in shape to that of the bandpass filter response shown in Fig. 9-2(c). Roll-off rate Sg Outside the passband, the rectangle-shaped idealized responses shown in Fig. 9-2 have infinite slopes, but of course, the actual responses have finite slopes. The steeper the slope, the more discriminating the filter is, and the closer it approaches the idealized response. Hence, the slope Sg outside the passband (called the gain roll-off rate) is an important attribute of the filter response.

L Zin2 R

+ _ VR

(b) Series RLC circuit Figure 9-4: Resonance occurs when the imaginary part of the input impedance is zero. For the RL circuit, Im [Zin1 ] = 0 when ω = 0 (dc), but for the RLC circuit, Im [Zin2 ] = 0 requires that ZL = −ZC or, equivalently, ω2 = 1/LC.

9-1.2

RC Circuit Example

To illustrate the transfer-function concept with a concrete example, let us consider the series RC circuit shown in Fig. 9-5(a). Voltage source Vs is designated as the input phasor, and on the output side, we have designated two voltage phasors, namely VR and VC . We now examine the frequency responses of the transfer functions corresponding to each of those two output voltages.

“book” — 2015/5/4 — 7:20 — page 505 — #6

9-1 THE TRANSFER FUNCTION

505

Vs

+ _ VC + _ VR

C

+ + _ −

~

R (a) RC circuit

ϕC

MC 1

0

1 = 0.707 √2

0

ωc

ω Lowpass phase

Lowpass magnitude

−π/4

0

ω

ωc

0

−π/2

(b) Magnitude and phase angle of HC(ω) = VC / Vs ϕR

MR 1

π/2

1 = 0.707 √2

Highpass magnitude

Highpass phase π/4

0

ωc

0

ω

ω

0

ωc 0 (c) Magnitude and phase angle of HR(ω) = VR / Vs Figure 9-5: Lowpass and highpass transfer functions.

Lowpass filter

The transfer function corresponding to VC is

Application of voltage division gives VC =

Vs ZC Vs /j ωC = . 1 R + ZC R + j ωC

(9.12)

HC (ω) =

VC 1 = , Vs 1 + j ωRC

(9.13)

“book” — 2015/5/4 — 7:20 — page 506 — #7

506

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

where we have multiplied the numerator and denominator of Eq. (9.12) by j ωC to simplify the form of the expression. In terms of its magnitude MC (ω) and phase angle φC (ω), the transfer function is given by HC (ω) = MC (ω) ej φC (ω) ,

C2

L

Vs

+

+ _

C1

R

Vout

_

(9.14)

with

Figure 9-6: Circuit of Example 9-1. MC (ω) = |HC (ω)| = √

1 1 + ω2 R 2 C 2

(9.15a)

and

Example 9-1: Resonant Frequency

φC (ω) = − tan

−1

(ωRC).

(9.15b)

Spectral plots for MC (ω) and φC (ω) are displayed in Fig. 9-5(b). It is clear from the plot of its magnitude that the expression given by Eq. (9.13) represents the transfer function of a lowpass filter with a dc gain factor M0 = 1. At dc, the capacitor acts like an open circuit—allowing no current to flow through the loop—with the obvious consequence that VC = Vs . At very high values of ω, the capacitor acts like a short circuit, in which case the voltage across it is approximately zero. Application of Eq. (9.6) allows us to determine the corner frequency ωc as follows MC2 (ωc ) =

1 1 = , 2 2 2 1 + ωc R C 2

(9.16)

For the circuit in Fig. 9-6, (a) obtain an expression for H(ω) = Vout √ /Vs and (b) show that H(ω) becomes purely real at ω0 = 1/ L(C1 + C2 ) . Solution: (a) Application of KCL and KVL leads to H(ω) =

RZC1 Vout , = Vs ZC1 ZC2 + ZL (ZC1 + ZC2 ) + R(ZC1 + ZL )

where ZL = j ωL, ZC1 = 1/j ωC1 , and ZC2 = 1/j ωC2 . After a few steps of algebra aimed at transforming the expression into a form whose denominator is purely real, we end up with H(ω) =

ω2 R 2 C22 (1 − ω2 LC1 ) + j ωRC2 [1 − ω2 L(C1 + C2 )] [1 − ω2 L(C1 + C2 )]2 + ω2 R 2 C22 (1 − ω2 LC1 )2

which leads to 1 . (9.17) RC Note that ωc is the inverse of the time constant τ = RC introduced in Chapter 5. ωc =

(b) At ω = ω0 = √

H(ω0 ) =

The output across R in Fig. 9-5(a) leads to VR j ωRC = . Vs 1 + j ωRC

and φR (ω) =

ωRC 1 + ω2 R 2 C 2

π − tan−1 (ωRC). 2

Their spectral plots are displayed in Fig. 9-5(c).

,

C1 + C2 . C2

(9.18) Concept Question 9-1: Is the transfer function of a circuit

always the same as its voltage gain? (See

The magnitude and phase angle of HR (ω) are given by MR (ω) = |HR (ω)| = √

L(C1 + C2 )

the imaginary part of the expression becomes equal to zero and the expression simplifies to

Highpass filter

HR (ω) =

1

(9.19a)

)

Concept Question 9-2: Is the gain factor M0 always the

peak value of M(ω)? (See

)

(9.19b) Concept Question 9-3: When is a circuit in a resonance

condition? (See

)

.

“book” — 2015/5/4 — 7:20 — page 507 — #8

9-2

SCALING

507

Concept Question 9-4: Why is the corner frequency also

called the half-power frequency? (See

)

Exercise 9-1: A series RL circuit is connected to a voltage source Vs . Obtain an expression for H(ω) = VR /Vs , where VR is the phasor voltage across R. Also, determine the corner frequency of H(ω). Answer: H(ω) = R/(R + j ωL), ωc = R/L. (See

)

9-2.1

Magnitude Scaling

The transfer function of a circuit is based on the impedances of its elements. If all impedances are multiplied (scaled) by the same magnitude scaling factor Km , the absolute level of the transfer function may or may not change, but its relative frequency response will remain the same. To distinguish between the prototype and scaled circuits, we shall: (a) Denote elements and impedances of the prototype circuit with unprimed symbols: ZR = R,

Zin

Answer: ω0 =

(9.20)

ZC =

and

(scaled circuit)

1 . j ωC  (9.21)

Magnitude scaling by a factor Km implies that: ZR = Km ZR ,

Figure E9.2



ZL = j ωL ,

C L

1 . j ωC

(b) Denote elements and impedances of the scaled circuit with primed symbols: ZR = R  ,

R

ZC =

and

(prototype circuit)

Exercise 9-2: Obtain an expression for the input

impedance of the circuit in Fig. E9.2 and then use it to determine the resonant frequency.

ZL = j ωL,

ZL = Km ZL ,

and

ZC = Km ZC , (9.22)

which translates into the relations R2

1 − 2 . (See LC L

R  = Km R,

)

L = Km L C =

C , Km

(9.23)

ω = ω .

9-2 Scaling

(magnitude scaling only) When designing a frequency filter, it often is convenient to start by designing a prototype model in which the elements have values on the order of ohms, henrys, and farads and then to scale the prototype circuit into a practical circuit that not only contains elements with realistic values but also provides the specified frequency response. A circuit can be scaled in magnitude, in frequency, or both.

Thus, resistor and inductor values scale by Km , but capacitor values scale by 1/Km . To illustrate with an example, consider the transfer function given by Eq. (9.18), HR (ω) =

j ωRC 1 + j ωRC

(9.24a)

j ωR  C  . 1 + j ωR  C 

(9.24b)

and its scaled version  Magnitude scaling changes the values of the elements in the circuit, but it does not modify its frequency response. Frequency scaling allows the designer to translate the frequency response into higher or lower frequency ranges while keeping the impedances of the circuit elements unchanged. 

HR (ω) =

Applying the recipe given by Eq. (9.23) leads to HR (ω) = HR (ω), which means that the frequency response remains unchanged.

“book” — 2015/5/4 — 7:20 — page 508 — #9

508

9-2.2

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

Frequency Scaling

To shift the profile of a transfer function along the ω axis by a frequency scaling factor Kf while keeping its relative shape the same, we can replace ω in the transfer function of the prototype circuit with ω = Kf ω and scale the element values so that their impedances remain unchanged. For an inductor, ZL = j ωL, so if ω is to be scaled up by Kf , L has to be scaled down by the same factor in order for ZL to stay the same. Hence, the impedance condition requires that 

R = R, L , Kf C C = , Kf

L1 = 3 H

Vs

C=

Prototype model

+

2 F 3

R=2Ω

Vout

_

(a) L1 = 3 mH

Vs

L =

+ _

L2 = 1 H

+ _

L2 = 1 mH 2 C = nF 3

Scaled version

+ R = 2 kΩ

Vout

_

(b) (9.25) Figure 9-7: Prototype and scaled circuits of Example 9-2.

ω = Kf ω. (frequency scaling only)

9-2.3

version with a cutoff frequency of ωc = 106 rad/s and a resistor value of 2 k.

Combined Magnitude and Frequency Scaling

To transfer the prototype circuit design into a realizable circuit, we often apply magnitude and frequency scaling simultaneously, in which case the relationships between the prototype and scaled circuits become

Km =

2k R = = 103 R 2

and

Kf =

ωc 106 = 106 . = ωc 1

Application of Eq. (9.26) leads to

R  = Km R, Km L, Kf 1 C = C, K m Kf

Solution: Based on the given information, the scaling factors are

Km 103 L1 = 6 × 3 = 3 mH, Kf 10 Km L2 = L2 = 1 mH, Kf L1 =

L =

(9.26)

ω = Kf ω. (magnitude and frequency scaling)

and C =

1 1 2 2 C= 3 × = nF. Km K f 10 × 106 3 3

The scaled circuit is displayed in Fig. 9-7(b). Example 9-2: Third-Order LP Filter

As discussed in Section 9-5, the order of a filter provides a measure of how steep its response is as a function of ω. The circuit in Fig. 9-7 is a prototype model of a third-order lowpass filter with a cutoff frequency of ωc = 1 rad/s. Develop a scaled

Concept Question 9-5: How is the scaling concept used in the design of resonant circuits and filters? (See ) Concept Question 9-6: What remains unchanged in (a) magnitude scaling alone and (b) frequency scaling alone? (See )

“book” — 2015/5/4 — 7:20 — page 509 — #10

TECHNOLOGY BRIEF 22: NOISE-CANCELLATION HEADPHONES

509

Technology Brief 22 Noise-Cancellation Headphones Noise-cancellation headphones are a class of devices that use active noise-control technology to reduce the level of environmental noise reaching a listener’s ear. They were invented by Amar Bose in 1978, based on concepts developed in the mid-20th century (most notably by Paul Lueg, who developed a system for cancelling noise in air ducts using loudspeakers). The primary advantage of such systems is the ability to selectively reduce noise without having to use heavy and expensive sound padding. Beyond hearing aids and the commercial headphones used by airline passengers, specialized active noise-control systems have been in use by pilots and heavy equipment operators for several decades. In small enclosed environments, active noise-control systems can employ microphone and speaker arrays to lower the amount of ambient noise experienced by the listener. Examples of this include noise-cancellation systems used to dampen engine noise in cockpits, active mufflers for industrial exhaust stacks, noise reduction around large fans and, recently, systems for reducing road and traffic noise in automobile interiors.

Active Noise Control In its most basic form, active noise control consists of measuring the sound levels at certain points in the environment and then using that data to emit noise from speakers whose frequency, phase shift, and amplitude are selected in order to cancel out the incoming environmental noise (Fig. TF22-1). In noisecancellation headphones, small microphones outside the headphones measure the incoming ambient noise, and the measured signal is then fed to circuitry that produces output noise in the headset that cancels out the ambient sound (Fig. TF22-2). The general phenomenon whereby one waveform is added to another to cancel it out is called destructive interference. The basic idea is to add a replica of the environmental noise signal, but shifted in phase by 180◦ , which is equivalent to multiplying the added signal waveform by (−1). Consider a vibrating wave traveling along a one-dimensional string (Fig. TF22-3), which is analogous to a sound pressure wave moving through the air. If we superimpose a second traveling wave onto the string (perhaps by waving the end up and down with a second hand), the two waves will overlap and the result will be the sum of the two individual waves (superposition). If we precisely time the second wave so

ENVIRONMENTAL SOUND

Microphone

Microphone

Control

Emitted sound Speaker

In specific region, emitted sound cancels the environmental sound Figure TF22-1: Active noise control. A basic active noise-control system uses a set of microphones to sense incoming ambient sound; the microphone signals are fed into control circuits which drive a set of speakers. The control circuit generates exactly the signals required to cancel out the incoming ambient sound in a specific region. that it is the exact mirror of the first (i.e., it is phase-shifted by 180◦ from the first wave), the two waves will cancel out exactly, and the string will not vibrate. This (in principle) is what active noise control aims to do, even though (in reality) the technology faces a number of limitations.

Limitations In order to truly cancel out all ambient noise, the emitted noise would have to exactly match the ambient noise in both space and time for all audible frequencies across a three-dimensional volume (such as the interior of a car or an airplane cabin). This is very difficult to accomplish in real environments. High frequencies are the hardest

“book” — 2015/5/4 — 7:20 — page 510 — #11

510

TECHNOLOGY BRIEF 22: NOISE-CANCELLATION HEADPHONES

Insulation attenuates high frequencies

Microphone

Speaker Control

ENVIRONMENTAL NOISE

Emitted sound cancels environmental sounds

Figure TF22-2: Noise-cancellation headphones employ the active noise-control principle to eliminate noise around the ear. Sound insulation is used commonly to remove high-frequency noise signals, while the active system removes low-frequency sound.

Traveling wave

Wave 1 y(t) = cos (ωt) Traveling wave

Wave 2

y(t) = cos (ωt + 180)

Wave 1 + Wave 2

No wave

FigureTF22-3: Destructive interference via the superposition of two waves. A string can be agitated so as to generate traveling waves along its length. In order to cancel out a wave traveling along the string (Wave 1), we simultaneously can generate a second wave of the same amplitude and frequency with a 180◦ phase shift (Wave 2). Because the vibration of the string will be the result of the two waves superimposed, the two waves cancel out, and no vibration occurs. This is analogous to what active control systems do: Microphones sense ambient waves that are then canceled out by emitting an appropriately phase-shifted wave from the speakers.

“book” — 2015/5/4 — 7:20 — page 511 — #12

TECHNOLOGY BRIEF 22: NOISE-CANCELLATION HEADPHONES

Ambient sound

Microphone

511

Speaker

Listener

Control

Input

Feedforward

Output

(a)

Feedback loop Speaker

Microphone

Control

Output

Input (b)

Figure TF22-4: Comparison between (a) an active noise-control system in a feedforward configuration, and (b) in a feedback configuration.

to match, because to correctly cancel them, the system would need to employ large arrays of microphones and speakers. Moreover, objects within the environment will reflect, absorb, and emit sound—further complicating the signals required to cancel the ambient sound. The situation is somewhat easier for headphones, as the area of interest is simply the user’s ear (a much smaller physical region). However, most commercial noise-cancellation headphones do not attempt to cancel high frequencies. Padding and passive layers are instead used to absorb the high frequencies and the system actively cancels out only the lower frequencies (e.g., the airplane engine hum). In general, noise cancellation only works well for sound that is periodic. Noise that is random or has very fast changes is very hard to mask, because the system cannot compute what the interfering signal should be instantaneously.

Feedforward versus Feedback Control Active noise-control systems provide an interesting comparison between feedback control (which we examined in Chapter 4) and feedforward control (Fig. TF22-4).

Consider again Fig. TF22-1. If the microphones of this system are positioned relatively far away from the speakers, they sample the incoming ambient sound signal and send it ahead to the control circuit, which then drives the speakers. There is no microphone at the speaker location and thus no way to measure the “output” of the system (i.e., there is no microphone that measures how well the system is canceling the sound at the listener). This is an example of feedforward control. If we were to move the microphones very close to the speakers (or, better yet right next to the listener), the microphones would continuously report how well the speakers were canceling the sound. If the control system is doing poorly, the microphones will detect some sound, and the control circuit can attempt to correct for this. In such a configuration, the system is operating with feedback control. Figure TF22-4 illustrates both of these control configurations. Some sophisticated active noise-control systems use both modes simultaneously: They have distant microphones as well as microphones near the listener. In general, feedforward systems are less practical to implement in consumer systems, and feedback systems tend to be less stable.

“book” — 2015/5/4 — 7:20 — page 512 — #13

512

CHAPTER 9

Exercise 9-3: Determine (a) Zin of the prototype circuit

shown in Fig. E9.3 at ω = 1 rad/s and (b) Zin of the same circuit after scaling it by Km = 1000 and Kf = 1000.

2Ix

R=2Ω

R=1Ω Ix

Zin

C=1F

Figure E9.3 Answer: (a) Zin = (1 − j ) , (b) Zin = (1 − j ) k.

(See

)

9-3 Bode Plots In the late 1930s, inventor Hendrik Bode (pronounced Bohdee) developed a graphical technique that has since become a standard tool for the analysis and design of resonant circuits, including filters, oscillators, and amplifiers.  Bode’s technique, which generates what we today call Bode plots or a Bode diagram, relies on using a logarithmic scale for ω and on expressing the magnitude of the transfer function in decibels (dB).  To make sure the reader is fully familiar with the properties of the dB operator, a quick review is in order.

9-3.1 The dB Scale The ratio of the power P relative to a reference power level P0 — such as the output power generated by an amplifier, relative to the input power supplied by the source—is called relative or normalized power. In many engineering applications, P /P0 may vary over several orders of magnitude when plotted against a specific variable of interest, such as the frequency ω of the circuit. The dB scale originally was introduced as a logarithmic conversion tool to facilitate the generation of plots involving relative power, but its use has since been expanded to other physical quantities. The dB operator is intended as a scale converter of relative quantities, such as P /P0 , rather than of P itself, but it still can be applied to P by setting P0 equal to

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

Table 9-1: Correspondence between power ratios in natural numbers and their dB values (left table) and between voltage or current ratios and their dB values (right table).

 V  V

       or  I   I  0 0

P P0

dB

10N 103 100 10 4 2 1 0.5 0.25 0.1 10−N

10N dB 30 dB 20 dB 10 dB ≈ 6 dB ≈ 3 dB 0 dB ≈ −3 dB ≈ −6 dB −10 dB −10N dB

dB

10N 103 100 10 4 2 1 0.5 0.25 0.1 10−N

20N dB 60 dB 40 dB 20 dB ≈ 12 dB ≈ 6 dB 0 dB ≈ −6 dB ≈ −12 dB −20 dB −20N dB

a specified value, such as 1 watt or 1 mwatt, so long as P is expressed in the same units as P0 . If G is defined as the power gain, G=

P , P0

(9.27)

then the corresponding gain in dB is defined as  G [dB] = 10 log G = 10 log

P P0

 (dB).

(9.28) The logarithm is in base 10. The dB scale converts a power ratio to its logarithmic value and then multiplies it by 10. Table 9-1 (left side) provides a listing of some values of G and the corresponding values of G [dB]. Note that when G varies across six orders of magnitude, from 10−3 to 103 , G [dB] varies from −30 dB to +30 dB. Also note that the dB value of 2 is ≈ +3 dB and the dB value of 0.5 is ≈ −3 dB. Even though the scale originally was applied to power ratios, it now is used to express voltage and current ratios as well. If P and P0 are the average powers absorbed by resistors of equal value and the corresponding phasor voltages across the resistors are V and V0 , respectively, then  G [dB] = 10 log

1 2 2 |V| /R 1 2 2 |V0 | /R



 = 20 log

 |V| . |V0 |

(9.29)

“book” — 2015/5/4 — 7:20 — page 513 — #14

9-3

BODE PLOTS

513

Similarly,

 G [dB] = 20 log

(b) The magnitude of H is given by



|I| . |I0 |

(9.30)

 Whereas the dB definition for power ratio includes a scaling factor of 10, the scaling factor for voltage and current is 20. 

M = |H| =

(ω/ωc ) (ω/ωc ) . =

|1 + j (ω/ωc )| 1 + (ω/ωc )2

(9.34)

Since H is a voltage ratio, the appropriate dB scaling factor is 20, so to find the power gain in dB, M [dB] = 20 log M

A useful property of the log operator is that the log of the product of two numbers is equal to the sum of their logs. That is if G = XY

G [dB] = X [dB] + Y [dB].

G [dB] = 10 log(XY ) = 10 log X+10 log Y = X [dB]+Y [dB]. By the same token, if: X Y

G [dB] = X [dB] − Y [dB].

(9.32)

Conversion of products and ratios into sums and differences will prove to be quite useful when constructing the frequency response of a resonant circuit. Example 9-3: RL Highpass Filter

For the series RL circuit shown in Fig. 9-8(a): (a) Obtain an expression for the transfer function H = Vout /Vs in terms of ω/ωc where ωc = R/L. (b) Determine the magnitude M [dB] = 20 log |H| and plot it as a function of ω on a log scale with ω expressed in units of ωc .

Solution: (a) Voltage division gives j ωLVs , R + j ωL

In the Bode-diagram terminology introduced later in Section 9-3.2, the components of M [dB] are called factors, so in the present case, M [dB] consists of two factors with the second one having a negative coefficient. A magnitude plot is displayed on semilog graph paper with the vertical axis in dB and the horizontal axis in (rad/s). If in the expression for M [dB], ω appears in a normalized format—as in (ω/ωc )—we may choose to express the horizontal axis in units of ωc . Figure 9-8(b) contains individual plots for each of the two factors comprising M [dB] as well as a plot for their sum. On semilog graph paper, the plot of log(ω/ωc ) is a straight line that crosses the ω axis at (ω/ωc ) = 1. This is because log 1 = 0. At (ω/ωc ) = 10, 20 log 10 = 20 dB. Hence;   1 20 log ω straight line with slope = 20 ωc dB/decade and ω axis crossing at ω/ωc = 1.  At ω/ωc = 10, 20 log(10) = 20 dB, at ω/ωc = 100, 20 log(100) = 40 dB, and so on. Hence, the slope is 20 dB/decade. 

properties: Low-Frequency Asymptote

which leads to Vout j ωL j (ω/ωc ) H= = = , Vs R + j ωL 1 + j (ω/ωc ) with ωc = R/L.

(9.35)

Note that a decade refers to a change by a factor of 10. Thus, the range from 1 to 10 is a decade, and so are the ranges from 3 to 30 and 10 to 100. 2 The second factor has a nonlinear plot, with the following

(c) Determine and plot the phase angle of H.

Vout =

= 20 log(ω/ωc ) − 10 log[1 + (ω/ωc )2 ]. 2 1

(9.31)

This result follows from

G=

= 20 log(ω/ωc ) − 20 log[1 + (ω/ωc )2 ]1/2

(9.33)

2a As (ω/ωc )

0,

−10 log 1 +



ω ωc

2 0.

“book” — 2015/5/4 — 7:20 — page 514 — #15

514

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

R

+

+ _

Vs

L

Vout

_

(a) RL circuit Degrees

dB 20 dB

90o

20 log(ω/ωc) 1

0 −3 dB

2a 0.1ωc

slope = 20 dB/ decade 10ωc

ωc

ω

M [dB] slope = −20 dB/ decade

ϕ(ω)

45o 0

0.1ωc

ωc

−10 log[1 +

ω

10ωc −45o 2

−90o

−tan−1(ω/ωc)

2b −20 dB

1

90o

(ω/ωc)2] 2

(b) Magnitude plot

(c) Phase plot

Figure 9-8: Magnitude and phase plots of H = Vout /Vs . High-Frequency Asymptote

(c) From Eq. (9.33), the phase angle of H is   ω . φ(ω) = 90◦ − tan−1 ωc 2 1

2b As (ω/ωc )

∞,  2 ω − 10 log 1 + ωc



ω − 20 log ωc

 .

The plot of M [dB] is obtained by graphically adding together the two plots of its individual factors (Fig. 9-8(b)). At low frequencies such that (ω/ωc  1), M [dB] is dominated by its first factor; at ω/ωc = 1, M [dB] = −3 dB; and at high frequencies (ω/ωc  1), M [dB] → 0, because its two factors cancel each other out. The overall profile (in red in Fig. 9-8(b)) is typical of the spectral response of a highpass filter with a cutoff frequency ωc .

(9.36)

The 90◦ component is contributed by j in the numerator and the second term is the phase angle of the denominator. The phase plot is displayed in Fig. 9-8(c). Concept Question 9-7: When is it helpful to use the dB

scale? (See

)

Concept Question 9-8: What is the scaling factor for power ratio? For current ratio? (See )

“book” — 2015/5/4 — 7:20 — page 515 — #16

9-3

BODE PLOTS

515

Exercise 9-4: Convert the following voltage ratios to dB:

(a) 20, (b) 0.03, (c) 6 × 106 . Answer: (a) 26.02 dB, (b) −30.46 dB, (c) 135.56 dB.

(See

)

Exercise 9-5: Convert the following dB values to voltage

ratios: (a) 36 dB, (b) −24 dB, (c) −0.5 dB. Answer: (a) 63.1, (b) 0.063, (c) 0.94. (See

9-3.2

)

Poles and Zeros

In polar coordinates, the transfer function H(ω) is composed of a magnitude M(ω) and a phase angle φ(ω), H(ω) = M(ω) ej φ(ω) .

(9.37)

For any circuit, the expression for H(ω) in general can be cast as the product of multiple factors A1 (ω) to An (ω), H(ω) = A1 (ω) A2 (ω) . . . An (ω).

(9.38)

Discussion of the functional forms of A1 to An will follow shortly, but to clarify what we mean by Eq. (9.38), let us consider the simple example of a transfer function given by H(ω) = 10

1 + j ω/ωz . 1 + j ω/ωp

the element values of the circuit. The quantity ωz (which has the same units as ω) is called a zero of H(ω), because it appears in a factor contained in the numerator of H(ω). Similarly, ωp is called a pole because it is part of a factor contained in the denominator of H(ω). If the numerator is a product of multiple factors, H(ω) will have multiple zeros—one associated with each factor (except for frequency independent factors, such as A1 = 10). A transfer function also may have multiple poles if the denominator of H(ω) is the product of multiple factors. Moreover, the factors may assume functional forms different from those given by A1 to A3 in Eq. (9.40). To analyze the frequency response of the circuit, we need to extract from Eq. (9.38) explicit expressions for the magnitude M(ω) and the phase angle φ(ω). For any two complex numbers, the phase angle of their product is equal to the sum of their individual phase angles. Application of this multiplication principle to Eq. (9.38) gives

(9.39)

φ(ω) = φA1 (ω) + φA2 (ω) + · · · + φAn (ω),

where φA1 (ω) to φAn (ω) are the phase angles of factors A1 to An , respectively. Transformation from a product form (as in Eq. (9.38)), into a sum (as in Eq. (9.41)) allows us to generate a phase plot for each factor separately and then add them together graphically—rather than having to deal with a single complicated expression all at once. The dB conversion introduced by Bode accomplishes a similar transformation for the magnitude M(ω). Application of the log property described by Eq. (9.31) leads to M [dB] = 20 log |H| = 20 log |A1 | + 20 log |A2 | + · · · + 20 log |An |

In this case, A1 = 10,

(9.40a)

A2 = 1 + j ω/ωz ,

(9.40b)

1 . 1 + j ω/ωp

(9.40c)

and A3 =

(9.41)

 The expression for H(ω) was structured intentionally into a form—called the standard form—in which the two terms involving ω each are written such that the real part is unity and the coefficient of ω in the imaginary part is defined as the reciprocal of an angular frequency (ωz or ωp ).  For the circuit represented by the transfer function given by Eq. (9.39), ωz and ωp are related to the circuit architecture and

= A1 [dB] + A2 [dB] + · · · + An [dB]

(9.42)

A1 [dB] = 20 log |A1 |,

(9.43)

where and a similar definition applies to the other factors. The transformations represented by Eqs. (9.41) and (9.42) constitute the basic framework for generating Bode diagrams. Our next step is to examine the possible functional forms that factors A1 to An may assume. Standard form refers to an arrangement in which factors A1 to An of Eq. (9.38) each can assume any one of only seven possible functional forms. We will examine the general character of each of these factors individually (as if it were the only component of H(ω)) by considering two types of plots: exact plots based on the exact expression for H(ω) and straightline approximations—called Bode plots—that are much easier to generate and yet provide reasonable accuracy. The symbol N (which we will call the order of a factor) is an integer equal to or greater than 1.

“book” — 2015/5/4 — 7:20 — page 516 — #17

516

CHAPTER 9

9-3.3

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS M [dB] = 20 log |K| φ = 0◦ if K > 0, or ±180◦ if K < 0 Same as exact plots; straight horizontal lines (Table 9-2)

Magnitude: Phase: Bode plots:

Functional Forms

Constant factor: H = K This is a frequency-independent constant that may be positive or negative.

Table 9-2: Bode straight-line approximations for magnitude and phase.

Factor

Bode Magnitude

Constant K



20 log K 0 dB

ω

0

Bode Phase 180 if K < 0 0 if K > 0 (90N)

Zero @ Origin ( jω)N

0 dB

Pole @ Origin

0 dB

1 1

( jω)−N

slope = 20N dB/decade ω ω slope = −20N dB/decade

ω



0

ω

0

ω (−90N)



(90N) Simple Zero (1 + jω/ωc)N Simple Pole

(

1 1 + jω/ωc

)

0 dB

0 dB

ωc ωc

N

slope = 20N dB/decade ω

ω

0 0.1ωc

ωc

10ωc

00.1ωc

ωc

10ωc

slope = −20N dB/decade

ω

ω

(−90N) (180N)

Quadratic Zero

slope = 40N dB/decade

[1 + j2ξω/ωc + ( jω/ωc)2]N 0 dB

Quadratic Pole

0 dB

1 [1 + j2ξω/ωc + ( jω/ωc)2]N

ωc ωc

ω

ω

0 0.1ωc

ωc

10ωc

00.1ωc

ωc

10ωc

slope = −40N dB/decade (−180N)

ω

ω

“book” — 2015/5/4 — 7:20 — page 517 — #18

9-3

BODE PLOTS

Zero @ origin: H = (j ω)N , (N = positive integer) The name of this factor reflects the fact that H → 0 as ω → 0. N is its order, so for example, (j ω)2 is called a second-order zero @ origin. Magnitude: Phase: Bode plots:

M [dB] = 20 log |(j ω)N | = 20N log ω φ = (90N )◦ Same as exact Magnitude: Straight line through ω = 1 with slope = 20N dB/decade Phase: Constant level (Table 9-2)

517

(1 + jω/ωc)N

dB 20N

Magnitude

Bode plot

Exact 3N 0 (a)

0.1ωc

10ωc

ωc

ω

Degrees

Pole @ origin: H = 1/(j ω)N This factor is called a pole because H → ∞ when ω → 0. The function 1/(j ω)3 is called a third-order pole @ origin, for example.    1   Magnitude: M [dB] = 20 log  (j ω)N  = −20N log ω Phase: φ = (−90N )◦ Bode plots: Same as exact (Table 9-2)

(b)

Magnitude and phase plots are identical to those of zero @ origin except for (−) sign in both cases.

Figure 9-9: Comparison of exact plots with the Bode straightline approximations for a simple zero with a corner frequency ωc .

Simple zero: H = (1 + j ω/ωc )N Standard form requires that the real part be 1 and the imaginary part be positive. The constant ωc is the corner frequency of the simple-zero factor, and N is its order.

Figure 9-9 provides a comparison between the Bode straightline approximation and the exact solution for both magnitude and phase. The corner frequency ωc gets its name from the Bode magnitude plot, which turns the corner at ωc .

Magnitude: M [dB] = 20 log |(1 + j ω/ωc )N | 2 = 10N

log[1 + (ω/ωc ) ] 0 dB, for ω/ωc  1 ≈ 20N log(ω/ωc ), for ω/ωc  1   ω −1 Phase: φ = N tan ωc

0, for ω/ωc  1 ≈ (90N )◦ , for ω/ωc  1 Bode plots: Straight-line approximation is different from exact; for magnitude, the maximum difference is 3N dB and it occurs at ω/ωc = 1 Magnitude: 0 dB horizontal line to ω = ωc , followed by straight line with slope = 20N dB/decade Phase: 0◦ horizontal line to ω = 0.1ωc ; straight line connecting coordinates [0.1ωc , 0] to [10ωc , (90N )◦ ]; followed by horizontal line (90N )◦

(90N) Phase

Exact Bode plot

(45N)

0 0.1ωc

ωc

10ωc

ω

Simple pole: H = 1/(1 + j ω/ωc )N Plots are mirror images (relative to ω axis) of those for the simple zero (Table 9-2). Quadratic zero: H = [1 + j 2ξ ω/ωc + (j ω/ωc )2 ]N N is the order of the quadratic zero, ωc is its corner frequency, and ξ is its damping factor. Magnitude: ⎧ ⎫  2 2  2 ⎬ ⎨ ω ω M [dB] = 10N log 1− + 4ξ 2 ⎩ ωc ωc ⎭

0 dB for ω/ωc  1, ≈ 40N log(ω/ωc ) for ω/ωc  1 Phase:

 2ξ(ω/ωc ) φ = N tan 1 − (ω/ωc )2

◦ for ω/ωc  1, 0 ≈ (180N )◦ for ω/ωc  1 −1



“book” — 2015/5/4 — 7:20 — page 518 — #19

518

CHAPTER 9

Bode plots:

Magnitude: Same as simple zero except at twice the slope Phase: Same as simple zero except at twice the slope and twice the level at ω/ωc  1

Figure 9-10 displays plots of the magnitude and phase of the quadratic-zero factor with N = 1 for three different values of the damping coefficient. Whereas the value of ξ has little influence on the shape of the plots when ω/ωc  1 or ω/ωc  1, it exercises significant influence when ω is in the neighborhood of ωc . In terms of the Bode straight-line approximation, the Bode plots (Table 9-2) for a quadratic factor of order N are identical to those for a simple factor of order 2N . )2 ]−N

Quadratic pole: H = [1 + j 2ξ ω/ωc + (j ω/ωc Plots are mirror images, relative to ω axis, as those for the quadratic zero.

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

[1 + j 2ξ ω/ωc + (j ω/ωc )2 ]

dB 40 30 Magnitude 20 Bode plot

10

Exact ξ = 0.5

0 − 10 −20

Exact ξ = 0.25

Exact ξ = 0.1 0.1ωc

10ωc

ωc

ω

(a) Degrees

9-3.4

General Observations

A few general observations about the Bode straight-line plots shown in Table 9-2 are in order. (1) For N = 1, the slope of the nonhorizontal Bode magnitude line (called gain roll-off rate) is 20 dB/decade for both the zero @ origin and simple-zero factors. The corresponding slope for their pole counterparts is −20 dB/ decade. (2) For N = 1, the slopes of the nonhorizontal Bode magnitude lines for the quadratic zero and quadratic pole factors are 40 dB/decade and −40 dB/decade, respectively. (3) The slopes of all Bode magnitude and phase lines are proportional to N . For example, the slope of the magnitude of a first-order simple-zero factor (1 + j ω/ωc ) is 20 dB/decade, so the slope of a third-order simple-zero factor (1 + j ω/ωc )3 is 60 dB/decade.

180 160 140 120

Phase Bode plot

100 80 60 40 20 0

Exact ξ = 0.1

Exact ξ = 0.5

Exact ξ = 0.25

0.1ωc

ωc

10ωc

ω

(b) Example 9-4: Bode Plots I

The voltage transfer function of a certain circuit is given by H(ω) =

(20 + j 4ω)2 . j 40ω(100 + j 2ω)

(a) Rearrange the expression into standard form. (b) Generate Bode plots for the magnitude and phase of H(ω).

Figure 9-10: Comparison of exact plots with Bode straightline approximations for a quadratic zero [1 + j 2ξ ω/ωc + (j ω/ωc )2 ].

Solution: (a) By factoring out 202 from the factor in the numerator and 100 from the factor in the denominator, we have H(ω) =

−j 0.1(1 + j ω/5)2 400(1 + j ω/5)2 = . j 4000ω(1 + j ω/50) ω(1 + j ω/50)

“book” — 2015/5/4 — 7:20 — page 519 — #20

9-3

BODE PLOTS

519

dB 3

40 dB Double zero at ωc = 5 rad/s

Magnitude

40 dB/decade

20 dB

Line segment 2 0 0.5

1

2

5

10

20

50

100

200

2 −20 dB 1

−20 log ω

M [dB]

−20 dB/decade

Line segment 1

ω (rad/s) 500 Simple pole at ωc = 50 rad/s −14dB −20 dB/decade

4

Line segment 3

(a) Degrees 2

180o

Double zero at ωc = 5 rad/s

Phase

90o ϕ(ω) 3

0 0.5

1

2

5

10

−90o

−90o

20

50

100

200

500

ω (rad/s)

Simple pole at ωc = 50 rad/s

1 (b)

Figure 9-11: Bode amplitude and phase plots for the transfer function of Example 9-4.

The corner frequency of the double-zero factor given by (1+j ω/5)2 is ωc1 = 5 rad/s, and similarly, the corner frequency of the simple-pole factor given by (1+j ω/50) is ωc2 = 50 rad/s. (b) M [dB] = 20 log |H| = 20 log 0.1 + 40 log |1 + j ω/5| − 20 log ω − 20 log |1 + j ω/50| = −20 dB + 40 log |1 + j ω/5| 3 1 − 20 log ω − 20 log |1 + j ω/50|. 2 4

The Bode line-approximations for the four terms constituting M [dB] and their sum are shown in Fig. 9-11(a). The sum is obtained by graphically adding the line-approximations corresponding to the four individual terms. An alternative method for generating the Bode magnitude plot is to start by plotting the line-approximation of the term with the lowest corner frequency, and then to move forward along the ω axis while sequentially changing the slope of the line as we encounter terms with higher corner frequencies. To illustrate the procedure, we labeled the three line segments of M [dB] in Fig. 9-11(a) as line segments 1, 2, 3, and 4. (1) The constant term is −20 dB (horizontal line with zero slope, labeled 1 in Fig. 9-11(a)).

“book” — 2015/5/4 — 7:20 — page 520 — #21

520

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

(2) The lowest-frequency term is the pole @ origin (1/ω). Its Bode line goes through 0 dB at ω = 1 rad/s and has a slope of −20 dB/decade. It is labeled 2 in Fig. 9-11(a)).

dB

(3) The combination of (1) and (2) generates line segment 1, which goes through ω = 1 at −20 dB.

30.4 dB

(4) The term with the next higher corner frequency is the double zero with ωc = 5 rad/s (labeled 3 in Fig. 9-11(a)). A double zero has a Bode line with a slope of +40 dB/decade. Hence, at ωc = 5 rad/s, we change the slope of line segment 1 by adding 40 dB/decade to its original slope of −20 dB/decade. This step generates line segment 2, with a slope of +20 dB/decade.

9.5 dB

(5) Line segment 2 continues until we encounter the corner frequency of the next term, namely the simple pole with ωc = 50 rad/s (labeled 4 in Fig. 9-11(a)). Adding a slope of −20 dB/decade leads to line segment 3, with a net slope of 0, and a constant level of −14 dB.

Double zero @ ωc = 3 rad/s M [dB]

20 dB

40 dB/decade 9.5 dB

0 0.1 0.3

1

3

10 30 100

ω (rad/s)

−40 dB/decade Quadratic pole at ωc = 10 rad/s

−20 dB

−40 dB (a) Degrees 180o

The phase of H(ω) is given by

Double zero at ωc = 3 rad/s

Phase

ω ω − tan−1 . 5 50 2 3

φ = −90◦ + 2 tan−1

1

Magnitude

40 dB

90o

ϕ(ω) 0

Bode plots for φ and its three components are shown in Fig. 9-11(b).

0.1 0.3

1

3

10 30 100

ω (rad/s)

−90o Quadratic pole at ωc = 10 rad/s

Example 9-5: Bode Plots II

−180o

Transfer function H(ω) is given by H(ω) =

(j 10ω + 30)2 . (300 − 3ω2 + j 90ω)

(a) Rearrange H(ω) into standard form. (b) Generate Bode plots for its magnitude and phase. Solution: (a) Upon reversing the order of the real and imaginary components in the numerator, factoring out 302 from it, and factoring out 300 from the denominator, we get 3(1 + j ω/3)2 , H(ω) = [1 + j 3ω/10 + (j ω/10)2 ] which consists of a constant factor K = 3, a zero factor with a corner frequency of 3 (rad/s), and a quadratic pole with a corner frequency of 10 rad/s.

(b)

Figure 9-12: Bode magnitude and phase plots for Example 9-5.

(b) M [dB] = 20 log |H| = 20 log 3 + 40 log |1 + j ω/3| − 20 log |1 + j 3ω/10 + (j ω/10)2 | = 9.5 dB + 40 log |1 + j ω/3| − 20 log |1 + j 3ω/10 + (j ω/10)2 |,   3ω/10 −1 −1 φ = 2 tan (ω/3) − tan . 1 − ω2 /100 Bode plots of M [dB] and φ are shown in Fig. 9-12. We note that M [dB] exhibits a highpass filter-like response.

“book” — 2015/5/4 — 7:20 — page 521 — #22

9-3

BODE PLOTS

521

M [dB] 30 dB −40 dB/decade 40 dB/decade

20 dB 19.1 dB −20 dB/decade 10 dB 20 dB/decade

0 0.5

1

2

3

5

10

20 30 50

100

200 300 500 1000

ω (rad/s)

Figure 9-13: Bode plot of bandreject filter of Example 9-6.

Exercise 9-6: Generate a Bode magnitude plot for the

transfer function

frequencies, but no gain to frequencies in the 10 to 50 (rad/s) range. Obtain the transfer function H(ω). Solution: The Bode plot consists of five segments. The first segment, corresponding to ω ≤ 1 rad/s, is generated by a pole @ the origin that goes through ω = 3 rad/s and has a slope of −40 dB/decade. The slope indicates that it is a double pole, so it must be given by   1 2 9 H1 = = 2. ω/3 ω

10(100 + j ω)(1000 + j ω) H= . (10 + j ω)(104 + j ω) Answer:

dB 20 dB

To verify the validity of our expression, let us convert it to dB as M1 [dB] = 20 log

0 10 (See

100

1000

104

ω (rad/s)

)

Example 9-6: Bandreject Filter

The Bode magnitude plot shown in Fig. 9-13 belongs to a bandreject filter that provides significant gain at low and high

9 = 20 log 9 − 40 log ω ω2 = 19.1 dB − 40 log ω.

At ω = 1 rad/s, M1 [dB] = 19.1 dB, which matches the figure. As we progress along the ω axis, the second segment has a slope of only −20 dB/decade, which means that a simple-zero factor with a corner frequency of 1 rad/s has come into play. Hence, H2 = (1 + j ω). At ω = 10 rad/s, the slope becomes zero, signifying the introduction of another simple-zero factor given by H3 = (1 + j ω/10).

“book” — 2015/5/4 — 7:20 — page 522 — #23

522

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

Similarly,

I

H4 = (1 + j ω/50) and H5 = (1 + j ω/200).

Vs

Hence,

+ _

H(ω) = H1 H2 H3 H4 H5 =

9(1 + j ω)(1 +

jω jω jω 10 )(1 + 50 )(1 + 200 ) . ω2

Since we are not given any information about the phase pattern of H(ω), the solution we obtained is correct within a multiplication factor of j N , which can accommodate j (for N = 1), −1 (for N = 2), −j (for N = 3), and 1 (for N = 4). The magnitude of H(ω) is the same, regardless of the value of N .

R

+ _ VR

L

+ _ VL

C

+ _ VC

+ VLC

_

Figure 9-14: Series RLC circuit.

9-4

Passive Filters

Concept Question 9-9: What does the term standard

Filters are of two types: passive and active.

form of a transfer function refer to, and what purpose does it serve? (See )

 Passive filters are resonant circuits that contain only passive elements, namely resistors, capacitors, and inductors. In contrast, active filters contain op amps, transistors, and/or other active devices, in addition to the passive elements. 

Concept Question 9-10: For which of the seven standard factors are the Bode plots identical to the exact plots and for which are they different? (See ) Concept Question 9-11: What is the gain roll-off rate?

(See

)

Passive and active filters are the subject of the next four sections. Any circuit that does not have a uniform frequency response is (by definition) a filter, simply because its output favors certain frequency ranges over others. Of particular interest to circuit designers are the four basic types of filters we introduced in Section 9-1. As we mentioned there, a filter transfer function is characterized by a number of attributes, including the following:

Exercise 9-7: Determine the functional form of the transfer function whose Bode magnitude plot is shown in Fig. E9.7, given that its phase angle at dc is 90◦ .

dB

1. The frequency ranges of its passband(s) and stopband(s).

40 dB

2. The gain factor M0 . 3. The gain roll-off rate Sg .

20 dB

0 2

20

500

5000

ω (rad/s)

Figure E9.7 Answer: H =

j (1 + j ω/2)(1 + j ω/500) . (See (1 + j ω/20)(1 + j ω/5000)

)

The objective of this section is to examine the basic properties of passive filters by analyzing their transfer functions. To that end, we use the series RLC circuit shown in Fig. 9-14, in which we have designated four voltage outputs, namely VR , VL , and VC across the individual elements, and VLC across the combination of L and C. We will examine the frequency responses of the transfer functions corresponding to all four output voltages.

“book” — 2015/5/4 — 7:20 — page 523 — #24

9-4

PASSIVE FILTERS

9-4.1

523

Bandpass Filter L

The current I flowing through the loop in Fig. 9-15(a) is given by j ωCVs  = I= , 1 (1 − ω2 LC) + j ωRC R + j ωL − ωC (9.44) where we multiplied the numerator and denominator by j ωC to simplify the form of the expression. The transfer function corresponding to VR is

+ R

ωRC (1 − ω2 LC)2

and

φBP(ω) = 90◦ − tan−1



+ ω2 R 2 C 2

 ωRC . 1 − ω2 LC

MBP 1 Idealized

(9.45)

0.707 Actual

, (9.46)

Bandpass (9.47)

According to the spectral plot displayed in Fig. 9-15(b), MBP goes to zero at both extremes of the frequency spectrum and exhibits a maximum across an intermediate range centered at ω0 . Hence, the circuit functions like a bandpass (BP) filter, allowing the transmission (through it) of signals whose angular frequencies are close to ω0 and discriminating against those with frequencies that are far away from ω0 . The general profile of MBP (ω) can be discerned by examining the circuit of Fig. 9-15(a) at specific values of ω. At ω = 0, the capacitor behaves like an open circuit, allowing no current to flow and no voltage to develop across R. At ω = ∞, it is the inductor that acts like an open circuit, again allowing no current to flow. In the intermediate frequency range when the value of ω is such that ωL = 1/ωC, the impedances of L and C cancel each other out, reducing the total impedance of the RLC circuit to R and the current to I = Vs /R. Consequently, VR = Vs , and HBP = 1. To note the significance of this specific condition, we call it the resonance condition, and we refer to the frequency at which it occurs as the resonant frequency ω0 , which is given by 1 . ω0 = √ LC

VR

_

(a) RLC circuit

where we added the subscript “BP” in anticipation of the fact that HBP (ω) is the transfer function of a bandpass filter. Its magnitude and phase angle are given by MBP (ω) = |HBP (ω)| =

I

+ _

Vs

Vs

VR RI j ωRC HBP (ω) = , = = Vs Vs (1 − ω2 LC) + j ωRC

C

(9.48)

0

0

ωc1 ω0 B

ωc2

2ω0

3ω0

2ω0

3ω0

ω

(b) MBP(ω) ϕBP 90o

0

ω0

ω

−90o (c) ϕBP(ω) Figure 9-15: Series RLC bandpass filter.

The phase plot in Fig. 9-15(c) conveys the fact that φBP is dominated by the phase of C at low frequencies and by the phase of L at high frequencies, and φBP = 0 at ω = ω0 .

“book” — 2015/5/4 — 7:20 — page 524 — #25

524

CHAPTER 9

At resonance, the energy in the circuit oscillates back and forth between L and C.

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

MBP [dB] 0

Filter bandwidth

ω0

0.5ω0

−3 −5

The bandwidth of the bandpass filter is defined as the frequency range extending between ωc1 and ωc2 , where ωc1 2 (ω) = 0.5 or and ωc2 are the values of ω at which MBP √ 2 is MBP (ω) = 1/ 2 = 0.707. As we will see shortly, MBP proportional to the power delivered to the resistor in the RLC circuit. At resonance, the power is at its maximum, and at ωc1 and ωc2 , the power delivered to R is equal to 1/2 of the maximum possible, which explains why ωc1 and ωc2 also are referred to as the half-power frequencies (or the −3 dB frequencies on a dB scale). Thus, @ ωc1 and ωc2 .

ω

Poor selectivity Q=2 Bandwidth B3

−10

Medium selectivity Q=5

−15

Bandwidth B2

−20

High selectivity Q = 20 Bandwidth B1

−25 −30

1 2 MBP (ω) = 2

2ω0

ω0

0.5ω0

(9.49)

2ω0

B1 B2

Upon inserting the expression for MBP (ω) given by Eq. (9.46) and carrying out several steps of algebra, we obtain the solutions

B3 Figure 9-16: Examples of bandpass-filter responses.

ωc1

R =− + 2L

ωc2

R = + 2L





R 2L

2 +

1 LC

(9.50a)

According to the foregoing discussion, the choice of values we make for R, L, and C will specify the overall shape of the transfer function completely, as well as its center frequency ω0 and bandwidth B.

and 



R 2L

2

1 + . LC

(9.50b)

 The quality factor of a circuit Q is an attribute commonly used to characterize the degree of selectivity of the circuit. A high Q circuit has a narrow bandwidth (relative to the center frequency) and high selectivity. 

(9.51)

Figure 9-16 displays frequency responses for three circuits, all with the same ω0 . The high-Q circuit exhibits a sharp response with a narrow bandwidth (relative to ω0 ), the medium-Q circuit has a broader pattern, and the low-Q circuit has a pattern with limited selectivity. For the bandpass-filter response, Q obviously is related to the ratio ω0 /B, but the formal definition of Q applies to any resonant circuit and is based on energy considerations, namely

The bandwidth then is given by

B = ωc2 − ωc1 =

R . L

It is worth noting that ω0 is equal to the geometric mean of ωc1 and ωc2 : ω0 =

√ ωc1 ωc2 .

Quality factor

(9.52)

 Q = 2π

 Wstor  , Wdiss ω=ω0

(9.53)

“book” — 2015/5/4 — 7:20 — page 525 — #26

9-4

PASSIVE FILTERS

525

where Wstor is the maximum energy that can be stored in the circuit at resonance (ω = ω0 ), and Wdiss is the energy dissipated by the circuit during a single period T . The stored energy is stored in L and C and the dissipated energy is dissipated in R. The factor 2π is an artificial multiplier introduced solely so that the expression for Q (that we will be deriving shortly) is simple in form and easy to remember. In the RLC circuit of Fig. 9-15(a), the source is represented by the phasor voltage Vs . In order to obtain expressions for the stored and dissipated energies at resonance, we need to (a) go back to the time domain and (b) specify ω as ω0 . For the sake of convenience and without loss of generality, we will assign the source the functional form υs (t) = V0 cos ω0 t

Vs = V0 .

(9.54)

At resonance, ω0 L = 1/ω0 C, so the expressions for the total impedance of the circuit and the current flowing through it simplify to √ (@ ω0 = 1/ LC) (9.55) Z = R + j ω0 L − j/ω0 C = R and

Vs V0 Vs = = . Z R R The time-domain current then is given by   V0 V0 j ω0 t = e cos ω0 t. i(t) = Re R R I=

V 2L 1 L iL2 (t) = 0 2 cos2 ω0 t 2 2R

and wC (t) =

1 1 C υC2 (t) = C 2 2 1 = C 2 =

 

V02 L 2R 2

1 C

T Wdiss = 0

T =

0

V02 π V02 cos2 ω0 t dt = . (9.60) R ω0 R

Upon substituting Eqs. (9.59) and (9.60) into Eq. (9.53), we obtain the result

Q=

ω0 L R

(bandpass filter).

(9.57)

(9.58a)

2

ω0 B

(ω/Qω0 ) {[1 − (ω/ω0 )2 ]2 + (ω/Qω0 )2 }1/2

and ◦

φBP (ω) = 90 − tan

(J).

(bandpass filter),

(9.58b)

−1



 (ω/ω0 ) . Q[1 − (ω/ω0 )2 ]

=

[cos2 ω0 t + sin2 ω0 t] =

.

(9.59)

(9.63a)

(9.63b)

Hence, the spectral response of the transfer function is specified completely by the combination of Q and ω0 . Also, in view of Eq. (9.61), the expressions given by Eq. (9.50) for the half-power frequencies ωc1 and ωc2 can be rewritten as

Wstor = wL (t) + wC (t) V02 L 2R 2

(9.62)

which is dimensionless. Thus, for a bandpass filter, Q is the inverse of the bandwidth B normalized to the center frequency ω0 . To highlight the role of Q, we can use the expressions for Q and ω0 to rewrite Eqs. (9.46) and (9.47) for the magnitude and phase angle of HBP (ω) in the forms MBP (ω) =

Even though both wL and wC vary with time, their sum is always a constant and equal to the maximum energy stored in the circuit,

V02 L 2R 2

(9.61)

Using the relation given by Eq. (9.51), the expression for the quality factor becomes

i dt

sin2 ω0 t

i R dt =

Q=

(J)

V0 sin ω0 t ω0 RC

2π/ω  0 2

0

2



pR dt

(9.56)

At any instant in time, the instantaneous energies stored in the inductor and the capacitor are given by wL (t) =

The energy dissipated by R during a single period is obtained by integrating the expression for the power pR over a period T = 1/f0 = 2π/ω0 so that

1 ωc1 =− + ω0 2Q

 1+

1 4Q2

(9.64a)

“book” — 2015/5/4 — 7:20 — page 526 — #27

526

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

Table 9-3: Attributes of series and parallel RLC bandpass circuits. C

L

+

+ Vs _

RLC Circuit

R H=

Transfer Function Resonant Frequency, ω0



Lower Half-Power Frequency, ωc1

C

VR Vs

H=

1



LC

R

V

_R

VR Is

1 LC

1 RC

ω0 ω0 L = B R  1 1 − ω0 + 1+ 2Q 4Q2

ω0 R = B ω0 L  1 1 − + 1+ ω0 2Q 4Q2

Upper Half-Power Frequency, ωc2

L

R L

Bandwidth, B Quality Factor, Q

V

_R

+

+ Is _

 1 + 2Q

1 ω0 1+ 4Q2



 1 + 2Q

1 ω0 1+ 4Q2

Notes: (1) The expression for Q of the series RLC circuit is the inverse of that for Q of the parallel circuit. (2) For Q ≥ 10, ωc1 ≈ ω0 − B2 , and ωc2 ≈ ω0 + B2 . and 1 ωc2 = + ω0 2Q

MBP [dB]

 1+

1 . 4Q2

(9.64b)

For a circuit with Q > 10, the expressions for ωc1 and ωc2 simplify to ωc1 ≈ ω0 −

B , 2

ωc2 ≈ ω0 +

B , 2

(if Q > 10) (9.65)

thereby forming a symmetrical bandpass centered at ω0 . Table 9-3 provides a summary of the salient features of the series RLC bandpass filter. For comparison, the table also includes the corresponding list for the parallel RLC circuit. Example 9-7: Filter Design

(a) Design a series RLC bandpass filter with a center frequency f0 = 1 MHz (Fig. 9-17) and a quality factor Q = 20, given that L = 0.1 mH.

0 −3 dB

ωa

ωc1

ω0

ωc2

ωb

ω

B 3 dB bandwidth

−10 dB 10 dB bandwidth ω0 = 2π

106

Figure 9-17: 10 dB bandwidth extends from ωa to ωb , corresponding to MBP (dB) = −10 dB.

(b) Determine the 10 dB bandwidth of the filter, which is defined as the bandwidth between frequencies at which the power level is 10 dB below the peak value. Solution: (a) Application of ω0 = 2πf0 = 2π × 106 = √

1 =√ LC 10−4 C 1

“book” — 2015/5/4 — 7:20 — page 527 — #28

9-4

PASSIVE FILTERS

527

leads to C = 0.25 nF. Solving Eq. (9.61) for R gives R=

C

L

ω0 L 2π × 106 × 10−4 = = 31.4 . Q 20

+ _

Vs

I1

(b) Voltage is proportional to MBP , and power is proportional 2 . The definition for power in dB is to MBP

C

L R

R

I2

+ _

Vo

(a) Two-stage circuit

2 P [dB] = 10 log P = 10 log MBP = 20 log MBP = MBP [dB].

M(ω) 1

We seek to find angular frequencies ωa and ωb corresponding to MBP [dB] = −10 dB (Fig. 9-17). If 20 log MBP = −10 dB,

0.8 0.707 0.6

MBP = 10−0.5 = 0.316.

0.4

it follows that The expression for MBP is given by Eq. (9.46) as MBP =

ωRC (1 − ω2 LC)2

+ ω2 R 2 C 2

and

0 9800

.

With MBP = 0.316, R = 31.4 , L = 10−4 C = 0.25 nF, solution of the expression yields ωa = 0.93 ω0

0.2 9900

ωc1 = 9963 rad/s

H, and

10000

10100

10200

ω (rad/s)

ωc2 = 10037 rad/s (b) M(ω)

ωb = 1.08. ω0

Figure 9-18: Two-stage RLC circuit of Example 9-8.

The corresponding bandwidth in Hz is B10 dB = (1.08 − 0.93) × 1 MHz = 0.15 MHz. and  −RI1 + I2 2R + j ωL +

Example 9-8: Two-Stage Bandpass Filter

Determine H(ω) = Vo /Vs for the two-stage BP-filter circuit shown in Fig. 9-18. If Q1 = ω0 L/R is the quality factor of a single stage alone, what is Q2 for the two stages in combination, given that R = 2 , L = 10 mH, and C = 1 μF?

ω0 = √

LC

=√

1 10−2 × 10−6

= 104 rad/s

and Q1 =

ω0 L 104 × 10−2 = = 50. R 2

The loop equations for mesh currents I1 and I2 are   1 + R − RI2 = 0 −Vs + I1 j ωL + j ωC

 = 0.

Simultaneous solution of the two equations leads to H(ω) =

Solution: For each stage alone, 1

1 j ωC

Vo Vs

=

ω2 R 2 C 2 ω2 R 2 C 2 − (1 − ω2 LC)2 − j 3ωRC(1 − ω2 LC)

=

ω2 R 2 C 2 [ω2 R 2 C 2 − (1 − ω2 LC)2 + j 3ωRC(1 − ω2 LC)] . [ω2 R 2 C 2 − (1 − ω2 LC)2 ]2 + 9ω2 R 2 C 2 (1 − ω2 LC)2

Resonance occurs when the imaginary part of H(ω) is zero, which is satisfied either √ when ω = 0 (which is a trivial resonance) or when ω = 1/ LC . Hence, the two-stage circuit has the same resonance frequency as a single-stage circuit. Using the specified values of R, L, and C, we can calculate the magnitude M(ω) = |H(ω)| and plot it as a function of ω.

“book” — 2015/5/4 — 7:20 — page 528 — #29

528

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

The result is displayed in Fig. 9-18(b). From the spectral plot, we have

C

R

ωc1 = 9963 rad/s,

+

+ _

Vs

I L

VL

_

ωc2 = 10037 rad/s, B2 = ωc2 − ωc1 = 10037 − 9963 = 74 rad/s,

(a) HHP = VL / Vs

and

MHP [dB]

ω0 104 Q2 = = = 135, B2 74

20

where B2 is the bandwidth of the two-stage BP-filter response. The two-stage combination increases the quality factor from 50 to 135.

Q = 10 (moderate resonance)

10 Q=2 (weak resonance)

Exercise 9-8: Show that for the parallel RLC circuit

0

shown in Fig. E9.8, the transfer-impedance transfer function HZ = VR /Is exhibits a bandpass-filter response.

+ _

Is

+ L

C

R

−10

Bode approximation

VR

_

−20

slope = 40 dB/decade

Figure E9.8

Passband

−30

Answer:

HZ =

Stopband

VR j ωL = . Is (1 − ω2 LC) + j ωL/R

The functional form of HZ (ω) is identical to that given by Eq. (9.45) for the series RLC bandpass filter. Moreover, √ both circuits resonate at ω = 1/ LC . (See C3 ) )

−40 0.1ω0

10ω0

ω0

ω

(b) Magnitude spectrum

Figure 9-19: Plots of MHP [dB] for Q = 2 (weak resonance) and Q = 10 (moderate resonance).

9-4.2

Highpass Filter

with magnitude and phase angle

At low frequencies, the capacitor C in the circuit of Fig. 9-19(a) acts like an open circuit, so VL across the inductor is essentially zero. Conversely, at high frequencies, the capacitor acts like a short circuit and the inductor acts like an open circuit. Consequently, VL ≈ Vs . This behavior constitutes a highpass filter. Transfer function HHP (ω), corresponding to VL in the circuit of Fig. 9-19(a), is given by VL j ωLI −ω2 LC = = HHP (ω) = 2 Vs Vs (1 − ω LC) + j ωRC

(9.66)

MHP (ω) = =

ω2 LC + ω2 R 2 C 2 ]1/2

[(1 − ω2 LC)2

(ω/ω0 )2 {[1 − (ω/ω0 )2 ]2 + (ω/Qω0 )2 }1/2

and φHP (ω) = 180◦ − tan−1 ◦

= 180 − tan

−1

 

ωRC 1 − ω2 LC

(9.67a)



 (ω/ω0 ) , Q[1 − (ω/ω0 )2 ]

(9.67b)

“book” — 2015/5/4 — 7:20 — page 529 — #30

9-4

PASSIVE FILTERS

529

where ω0 and Q are defined by Eqs. (9.48) and (9.61), respectively. Figure 9-19(b) displays logarithmic plots of MHP [dB] for two values of Q. Because MHP (ω) has a quadratic zero, its slope in the stopband is 40 dB/decade. Exercise 9-9: How should R be related to L and C so that

the denominator of Eq. (9.66) becomes a simple pole of order 2? What will the value of Q be in that case? √ Answer: R = 2 L/C , Q = 1/2. (See C3 ) )

L

R Vs

9-4.3

The voltage across the capacitor in Fig. 9-20(a) generates a lowpass-filter transfer function given by VC (1/j ωC)I 1 = = , 2 Vs Vs (1 − ω LC) + j ωRC (9.68) with magnitude and phase angle given by HLP (ω) =

1 + ω2 R 2 C 2 ]1/2 1 = , 2 2 {[1 − (ω/ω0 ) ] + (ω/Qω0 )2 }1/2

MLP (ω) =

I

+

+

~+−_

C

Lowpass Filter

VC

_

and

(a) HLP = VC / Vs

φLP (ω) = − tan

MLP [dB]

10

Q = 10 (moderate resonance)

0

Q=2 (weak resonance)

−20

 

ωRC 1 − ω2 LC

(9.69a)



 (ω/ω0 ) . Q[1 − (ω/ω0 )2 ]

(9.69b)

The spectral plots of MLP [dB] shown in Fig. 9-20(b) are mirror images of the highpass filter plots displayed in Fig. 9-19(b).

9-4.4

Bandreject Filter

The output voltage across the combination of L and C in Fig. 9-21(a) generates a bandreject filter transfer function and is equal to Vs − VR :

Bode approximation

HBR (ω) =

slope = −40 dB/decade Stopband

−30 Passband −40 0.1ω0

−1

= − tan−1

20

−10

[(1 − ω2 LC)2

ω0 (b) Magnitude spectrum Figure 9-20: RLC lowpass filter.

10ω0

ω

VL + VC Vs − VR = = 1 − HBP (ω), Vs Vs

(9.70)

where HBP (ω) is the bandpass filter transfer function given by Eq. (9.45). The spectral response of HBP passes all frequencies except for an intermediate band centered at ω0 , as shown in Fig. 9-21(b). The width of the stopband is determined by the values of ω0 and Q. Exercise 9-10: Is MBR = 1 − MBP ? Answer: No, because MBR = |HBR| = |1 − HBP| = 1 − |HBP| = 1 − MBP. (See )

“book” — 2015/5/4 — 7:20 — page 530 — #31

530

CHAPTER 9

Vs

the years to define the order of a filter, so to avoid ambiguity, we adopt the following definition.

I

R

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

+

+ _

L C

 The order of a filter is equal to the absolute value of the highest power of ω in its transfer function when ω is in the filter’s stopband(s). 

VLC

_

(a) HBR = VLC / Vs

Let us examine this definition for three circuit configurations.

9-5.1 Bandstop for Q = 10

First-Order Lowpass RC Filter

The transfer function of the RC circuit shown in Fig. 9-22(a) is given by

Bandstop for Q = 2 H1 (ω) =

MBR [dB] 0 −3 −5

ω0

0.5ω0

2ω0

ω (rad/s)

Q = 10 Q=2

−10

VC 1/j ωC = Vs R + 1/j ωC 1 = 1 + j ωRC 1 = 1 + j ω/ωc1

(first-order),

(9.71)

where we multiplied both the numerator and denominator by j ωC so as to rearrange the expression into the standard form we discussed in Section 9-3.2. The expression given by Eq. (9.71) is a simple pole with a corner frequency given by

−15 −20

Bandreject

ωc1 =

−25 −30

0.5ω0

ω0

2ω0

(b) Spectral response Figure 9-21: Bandreject filter.

9-5 Filter Order In Section 9-3, we associated the term order with the power of ω, so a factor given by (1 + j ω/ωc )2 was called a secondorder zero because the highest power of ω in the expression is 2. Similarly, (1 + j ω/ωc )−2 was called a second-order pole because the highest power of ω is also 2, but the expression appears in the denominator. The term order also is used to describe the overall filter response, which may be composed of the product of several zero and pole factors—each with its own order. Multiple, different characterizations have been used over

1 RC

(RC filter).

(9.72)

It is evident from the expression given by Eq. (9.71) that the highest order of ω is 1, and therefore the RC circuit is a firstorder filter. Strict application of the definition for the order of a filter requires that we evaluate the power of ω when ω is in the stopband of the filter. In the present case, the stopband covers the range ω ≥ ωc1 . When ω is well into the stopband (ω/ωc1  1), Eq. (9.71) simplifies to H1 (ω) ≈

−j ωc1 ω

(for ω/ωc1  1),

(9.73)

which confirms the earlier conclusion that the RC circuit is first-order.  A circuit containing a single reactive element (capacitor or inductor) generates a first-order transfer function. Generally speaking, the order of a filter depends on the number of reactive elements contained in the circuit. The order may be smaller or equal to the number of reactive elements, but not greater. 

“book” — 2015/5/4 — 7:20 — page 531 — #32

9-5

FILTER ORDER

531

R

+

+ + _ −

~

Vs (a)

C

Vs

VC

_

+

~

C

VC

_

Second-order filter M2 [dB]

M1 [dB] 0.1ωc1

10ωc1

ωc1

100ωc1

Bode

ω

−20 dB

0 −3 dB

0.1ωc2

10ωc2

ωc2

100ωc2

Bode

ω

−20 dB Sg = −20 dB/decade

Sg = −40 dB/decade

−40 dB (b)

I

+ + _ −

(c)

First-order filter

0 −3 dB

L

R

−40 dB Response of first-order filter

(d)

Response of second-order filter

Figure 9-22: Comparison of magnitude responses of the first-order RC filter and the second-order RLC filter. The corner frequencies are given by ωc1 = 1/RC and ωc2 = 1.28/RC.

The magnitude of H1 (ω) is given by M1 = |H1 (ω)| =

1 1 . (9.74) =

|1 + j ω/ωc1 | 1 + (ω/ωc1 )2

At ω = ωc1 , M1 (ωc1 ) = √

1 1+1

= 0.707.

(9.75)

When expressed in dB, M1 becomes

 For a first-order filter, Sg = −20 dB/decade. To achieve a faster rate of decay, second- or higher-order filters are called for.  Example 9-9: Filter Transmission Spectrum

M1 [dB] = 20 log M1 = −10 log[1 + (ω/ωc1 )2 ] ⎧ ⎪ @ ω = 0, ⎨0 dB = −3 dB @ ω/ωc1 = 1, ⎪ ⎩ −20 log(ω/ωc1 ) @ ω/ωc1  1.

steady-state value of −20 dB/decade at much greater values of ω. As noted earlier, the steepness (slope) of the transfer function after it has transitioned from its passband to its stopband is called its gain roll-off rate Sg .

(9.76)

On the semilog scale of Fig. 9-22(b), M1 [dB] starts out at 0 dB—corresponding to M1 = 1 in natural units—decreases to −3 dB at ω = ωc1 , and then its slope accelerates towards a

An RC lowpass filter uses a capacitor C = 10 μF. (a) Specify R so that ωc1 = 1 krad/s. (b) The filter is considered acceptably transparent to a signal if the signal’s voltage amplitude is reduced by no more than 12 dB as it passes through the filter. What is the filter’s transmission spectrum according to this criterion? Solution: (a) Application of Eq. (9.72) leads to R=

1 ωc1 C

=

103

1 = 100 . × 10−5

“book” — 2015/5/4 — 7:20 — page 532 — #33

532

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS L, and C. By replacing the minus sign in Eq. (9.77) with j 2 and selecting R such that

(b) If M1 [dB] = 20 log M1 = −12 dB,

 R=2

then log M1 = −

12 = −0.6 20

L C

(no-ripple condition),

(9.79)

the expression given by Eq. (9.77) can be converted into a perfect square:

and

Equating this value of M1 to the expression given by Eq. (9.74) leads to 1 M1 =

= 0.25, 1 + (ω/ωc1 )2 which yields the solution ω = 3.87 ωc1

or

1 √ 1 + j 2 ω2 LC + j 2 LC 1 = (second order). √ (1 + j ω LC)2

H2 (ω) =

M1 = 10−0.6 = 0.25.

ω = 3.87 krad/s.

The constraint given by Eq. (9.79) allowed us to convert H2 (ω) from a quadratic pole into a simple pole of second-order. Its magnitude response is displayed in Fig. 9-22(d). ) is The corner frequency of the RLC lowpass filter (ωc2√ determined by setting the magnitude of H2 (ω) equal to 1/ 2. Thus,

Hence, the transmission spectrum of the filter extends from 0 to 3.87 krad/s or equivalently from 0 to 616 Hz.

9-5.2

For the RLC circuit shown in Fig. 9-22(c), we determined in Section 9-4.3 that its transfer function is given by Eq. (9.68) as VC 1 = Vs (1 − ω2 LC) + j ωRC

|H2 (ωc2 )| =

1 1 =√ , 2 1 + ωc2 LC 2

which leads to

Second-Order Lowpass Filter

H2 (ω) =

(9.80)

(RLC filter).

(9.77) The magnitude spectrum of the RLC lowpass filter was presented earlier in Fig. 9-20(b), where it was observed that the response may exhibit√a resonance phenomenon in the neighborhood of ω0 = 1/ LC, and that it decays with Sg = −40 dB/decade in the stopband (ω ≥ ω0 ). This is consistent with the fact that the RLC circuit generates a secondorder lowpass filter when the output voltage is taken across the capacitor. In terms of our definition for the order of a filter in the stopband (ω2  1/LC), Eq. (9.77) reduces to

ωc2

√ 1/2 2−1 0.64 . = =√ LC LC

(9.81)

From Eq. (9.79), L = R 2 C/4. When used in Eq. (9.81), the expression for ωc2 becomes

ωc2 =

1.28 RC

(RLC filter).

(9.82)

The foregoing analysis warrants the following observations:

(9.78)

1. The RC lowpass filter is first-order, its corner frequency is ωc1 = 1/RC, and its gain roll-off rate is Sg = −20 dB/ decade.

which assumes the form of a second-order pole. The ripple-like effect exhibited by the RLC filter in Fig. 9-20 can be avoided through a judicious choice of the values of R,

2. By adding a series inductor whose value is specified by L = R 2 C/4, the filter becomes second-order, its corner frequency shifts upward to 1.28/RC, and its slope becomes twice as steep.

H2 (ω) ≈

−1 ω2 LC

(for ω  ω0 ),

“book” — 2015/5/4 — 7:20 — page 533 — #34

TECHNOLOGY BRIEF 23: SPECTRAL AND SPATIAL FILTERING

533

Technology Brief 23 Spectral and Spatial Filtering Filtering is applied in electrical systems, sound systems, optical systems, mechanical systems, and more. It is so ubiquitous that we often are unaware of the filtering that occurs in every day life. This Technology Brief provides an overview of how filters are applied in sight and photography, and how this affects what we see. There are two major types of filters in what we see—spectral filters that control color and spatial filters that control smoothing and edges.

Spectral Filters Our eye is, itself, a spectral filter. The colors we can see range in wavelength from about 750 nm (400 THz, red) to 400 nm (750 THz, violet). Other frequencies, such as the ultraviolet (UV) and infrared (IR), are not detectable by the human eye.Thus, our eye is a spectral bandpass filter with a bandwidth of about 350 THz. But we do not see all colors equally well. Our eye’s frequency response is nonlinear. Figure TF23-1 shows the relative response of the eye to

Figure TF23-2: The eyes are a nonlinear bandpass filter. Greenish yellow provides high visibility in both dim and bright conditions because of the nonlinear sensitivity of our eyes to this color. (Credit: Agoora.co.uk.)

different colors of light. The eye is more sensitive to colors in the yellow-green region than to red and blue colors.This is why neon-yellow clothing provides high visibility under both dim and bright conditions (Fig. TF23-2).

Visible light

Infrared

Ultraviolet

Near

Near A

B C

Far (vacuum)

nm

nm

nm

nm

nm

nm

nm

nm

μm

nm 0 20 nm 0 28 m n 0

32

0

40

0

45

0

50

0

55

0

60

0

65

0

70

0

75

4

μm

m m

1.

3

1

Figure TF23-1: Visible, infrared, and ultraviolet parts of the electromagnetic spectrum.

“book” — 2015/5/4 — 7:20 — page 534 — #35

534

TECHNOLOGY BRIEF 23: SPECTRAL AND SPATIAL FILTERING

Figure TF23-3: Field of view (spatial filter) for humans, dogs, horses, and birds. (Courtesy of Wordpress.com.) Our eyes are also a spatial filter. Humans can see across an angular range of about 180◦ , including a narrow range where we focus and see colors, and a much wider peripheral field of view where vision is not as clear and colors are more limited (Fig. TF23-3). The field of view is controlled mainly by the placement of the eyes on the head. Some birds have 360◦ fields of view. Some animals (horses, for instance) do not see directly in front of them,

but have broader fields of view that let them see along both of their sides and even almost behind them. Prey animals tend to have larger fields of view than predators, whose eyes have more focus in front of them. Tinted lenses and photo editing are spectral filters that can selectively filter out various parts of the optical spectrum (Fig. TF23-4), acting as band pass or band reject filters. Gray tinted lenses are most common,

(a) Grey filter

(a) Amber filter

(c) Rose filter

(d) Blue filter

Figure TF23-4: Color filters. (Credit: SunglassWarehouse.com and Krista Ward.)

“book” — 2015/5/4 — 7:20 — page 535 — #36

TECHNOLOGY BRIEF 23: SPECTRAL AND SPATIAL FILTERING

535

Normal vision (92%) Deuteranomaly (2.7%) Protanomaly (0.66%) Protanopia (0.59%) Deuteranopia (0.56%) Tritanopia (0.016%) Tritanomaly (0.01%) Achromatopsia ( ωHP)

ωLP

ω

ωHP

(b) Bandreject filter

Figure 9-25: (a) In-series cascade of a lowpass and a highpass filter generates a bandpass filter; (b) in-parallel cascading generates a bandreject filter.

and  M2 [dB] = 20 log

103



1 + (ω/105 )2

= 60 dB − 20 log[1 + (ω/105 )2 ].

Exercise 9-14: What are the values of the corner

frequencies associated with M1 , M2 , and M3 of Example 9-10? Answer: ωc1 = 105 rad/s, ωc2 = 0.64ωc1 = 6.4 × 104

rad/s, ωc3 = 0.51ωc1 = 5.1 × 104 rad/s. (See

)

Similarly, M3 [dB] = 60 dB − 30 log[1 + (ω/105 )2 ]. The three-stage process is shown in Fig. 9-26(b) in blockdiagram form, and spectral plots of M1 [dB], M2 [dB], and M3 [dB] are displayed in Fig. 9-26(c). We note that the gain roll-off rate Sg is −20 dB for M1 [dB], −40 dB for M2 [dB], and −60 dB for M3 [dB]. We also note that the −3 dB corner frequencies are not the same for the three stages. Concept Question 9-16: Why is it more practical to

cascade multiple stages of active filters than to cascade multiple stages of passive filters? (See ) Concept Question 9-17: What determines the gain factors of the highpass and lowpass op-amp filters? (See )

Analogy to AND and OR gates An AND logic gate has two inputs (whose logic states can each be either 0 or 1) and one output. Its output state is 1 if and only if both input states are 1. Otherwise, its output state is zero. Hence, the output of an AND gate is equal to the product of its input states. The cascaded bandpass filter diagrammed in Fig. 9-25(a) is analogous to an AND gate. The filter consists of an idealized lowpass filter with cutoff frequency ωLP connected in series with an idealized highpass filter with cutoff frequency ωHP . The frequency of the input signal has to be in the passband of both filters in order for the signal to make it to the output. The filter passband is defined by the bandwidth extending from ωHP to ωLP , as illustrated in Fig. 9-25(a). The combination of the two filters when cascaded in series is equivalent to an AND gate, because the final output is proportional to the product of their transfer functions, HLP HHP . The process is illustrated through Example 9-11.

“book” — 2015/5/4 — 7:20 — page 540 — #41

540

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

1 nF

1 nF

10 kΩ 10 Ω

Vs

10 kΩ

_

V1 10 kΩ

10 kΩ

_

+

+ _

1 nF

V2 10 kΩ

+

_ +

+ V3

_

(a) Circuit diagram

Vs

Lowpass ωLP = 105 rad/s G = −103

Lowpass ωLP = 105 rad/s G = −1

V1

V2

Lowpass ωLP = 105 rad/s G = −1

V3

(b) Block diagram dB

60 57 M1 [dB] −20 dB/decade (1st order)

50 40

M2 [dB] −40 dB/decade (2nd order)

30 20

M3 [dB] −60 dB/decade (3rd order)

10 0 104

ωc3

105 ωc2

ωc1

106

107

ω (rad/s)

(c) Transfer function plots Figure 9-26: Three-stage lowpass filter and corresponding transfer functions.

In contrast, the bandreject filter (Fig. 9-25(b)) is analogous to an OR gate, for which the state of its output is 1 if either one or both of its inputs has a state of 1. The cascade configuration of the bandreject filter consists of a lowpass filter connected in

parallel with a highpass filter—thereby offering the input signal to pass through either or both of them—and then their outputs are added by a summing amplifier. Example 9-12 provides more details.

“book” — 2015/5/4 — 7:20 — page 541 — #42

9-7

CASCADED ACTIVE FILTERS

541

M [dB] 35

29.8 Rf1 Rs1 Vs

_

Rs2

+

+ _

M0 [dB]

32.8

Cf1

Lowpass

Cs2

3 dB

30

Rf2

_ + Highpass

25

+ Vout

_

20

15 103

ωc1

104 ω0

ωc2

105

ω (rad/s)

(b) M [dB]

(a) Two-stage bandpass filter Figure 9-27: Active bandpass filter of Example 9-11.

Example 9-11: Bandpass Filter

The block diagram shown in Fig. 9-27(a) is a two-stage bandpass filter with the following elements: Rs1 = 1 k, Rf1 = 10 k, Cf1 = 9 nF, Rs2 = 12 k, Cs2 = 9 nF, and Rf2 = 96 k. Determine and plot the magnitude of the transfer function and obtain the values of ω0 , ωc1 , ωc2 , B, and Q. Solution: The first stage is a lowpass filter with a transfer function given by Eq. (9.88) as   1 HLP (ω) = GLP 1 + j ω/ωLP with GLP

Rf 104 = − 1 = − 3 = −10 Rs1 10

and ωLP =

1 1 = 4 = 11.11 krad/s. Rf1 Cf1 10 × 9 × 10−9

The transfer function of the highpass filter in the second stage is characterized by Eq. (9.91) as   j ω/ωHP HHP (ω) = GHP 1 + j ω/ωHP

with GHP = −

Rf2 96 × 103 =− = −8 Rs2 12 × 103

and ωHP =

1 1 = = 9.26 krad/s. Rs2 Cs2 12 × 103 × 9 × 10−9

The combined transfer function is then given by H(ω) = HLP HHP   j ω/ωHP = G1 G2 , (1 + j ω/ωLP )(1 + j ω/ωHP )

(9.93)

and its magnitude is M = |H(ω)|     80ω/(9.26 × 103 )  =  3 3 [1 + j ω/(11.11 × 10 )][1 + j ω/(9.26 × 10 )]  =

80ω/(9.26 × 103 ) . {[1 + (ω/11.11 × 103 )2 ][1 + (ω/9.26 × 103 )2 ]}1/2 (9.94)

The angular frequency ω0 of a bandpass filter is defined as the frequency at which the transfer function is a maximum.

“book” — 2015/5/4 — 7:20 — page 542 — #43

542

CHAPTER 9

For a high-Q filter, ω0 is approximately midway between the lower and upper cutoff frequencies ωc1 and ωc2 , but ω0 may be significantly closer to ω1 or ω2 if Q is not very large. We do not yet know the value of Q for the present filter, but we can generate an approximate estimate from knowledge of the values of ωLP and ωHP . The estimated bandwidth is B (est) = ωLP − ωHP = (11.11 − 9.26)k = 1.85 krad/s, and if we assume that ω0 is midway between ωHP and ωLP , then   1.85 ω0 (est) = 9.26 + k = 10.185 krad/s 2

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS (c) gain roll-off rate = −40 dB/decade along both boundaries of the bandstop. Solution: The specified roll-off rate requires cascading two identical lowpass filters and two identical highpass filters. To minimize performance variations among identical pairs, identical resistors will be used in all four units (Fig. 9-28(b)), which means that they all will have unity gain. The overall gain of −50 will be provided by the summing amplifier. Somewhat arbitrarily, we select R = 1 k, and the value of Rf is specified by the gain of the summing amplifier as G = −50 = −

and Q (est) =

ω0 (est) 10.185 = = 5.5. B (est) 1.85

Since Q is not greater than 10, the estimated values of B, ω0 , and Q are not likely to be very accurate, so we should return to the expression for M given by Eq. (9.94) and use it to determine the exact values of ω0 , ωc1 , and ωc2 . We can do so by calculating (or plotting) M as a function of ω to identify: (a) M0 and ω0 , the maximum value of M(ω) and the corresponding value of ω at which it occurs, respectively, and √ (b) ωc1 and ωc2 , the corner frequencies at which M = M0 / 2 (or −3 dB below the peak on a dB scale). According to the spectral plot of M [dB] shown in Fig. 9-27(b), M0 [dB] = 32.8 dB, ωc1 = 4.19 krad/s,

ω0 (exact) = 10.14 krad/s, and

ωc2 = 24.56 krad/s.

Hence, B (exact) = ωc2 − ωc1 = 20.37 krad/s and Q (exact) =

ω0 10.14 = ≈ 0.51. B 20.37

The obvious conclusion is that our estimated values for B and Q are way off in comparison with their exact counterparts. We assumed that the corner frequencies ωLP and ωHP associated with functions HLP and HHP , respectively, are good estimates of the corner frequencies ωc1 and ωc2 of the product of the two functions. That was obviously a poor assumption. Example 9-12: Bandreject Filter

Design a bandreject filter with the specifications: (a) Gain = −50, (b) bandstop extends from 20 kHz to 40 kHz, and

Rf R

Rf = 50 k.

The transfer function of the bandreject filter is given by H(ω) = G[H2LP + H2HP ]  2  2 j ωRCHP 1 . + = −50 1 + j ωRCLP 1 + j ωRCHP Next, we need to specify values for CLP and CHP . As an approximation, we assume that over the passband of the lowpass filters, the highpass filters exercise minimal impact, and vice versa. This allows us to deal with the transfer functions of the lowpass and highpass filters separately. The magnitude of the transfer function of the two cascaded lowpass filters is MLP =

|H2LP |

 2   1 1   = . = 2 2  1 + j ωRCLP  1 + ω R 2 CLP

The specifications call for a lower corner frequency of 20 kHz or, equivalently, ω1 = 2π × 2 × 104 = 4π × 104 rad/s. Setting √ MLP = 1/ 2 , R = 1 k, and ω = ω1 leads to CLP ≈ 5 nF. A similar analysis for the highpass filter chain with f2 = 40 kHz (or ω2 = 8π × 104 rad/s) leads to 2 ω2 R 2 CHP 2 1 + ω2 R 2 CHP

 1  = √  , 2 at ω=ω2

which provides the solution CHP ≈ 6 nF. The spectrum of M [dB] = 20 log |H| is displayed in Fig. 9-28(c).

“book” — 2015/5/4 — 7:20 — page 543 — #44

9-7

CASCADED ACTIVE FILTERS

Exercise 9-15: The bandreject filter of Example 9-12 uses

543



Answer:

H(ω) = 50

two lowpass-filter stages and two highpass-filter stages. If three stages of each were used instead, what would the expression for H(ω) be in that case?

1 1 + j ω/4π × 104

 +

(See

)

3

j ω/8π × 104 1 + j ω/8π × 104

3 .

“book” — 2015/5/4 — 7:20 — page 544 — #45

544

TECHNOLOGY BRIEF 24: ELECTRICAL ENGINEERING AND THE AUDIOPHILE

Technology Brief 24 Electrical Engineering and the Audiophile The reproduction of high-quality music with sufficient fidelity to sound like a live performance in one’s living room was one of the technological hallmarks of the 20th century. In these days of iPods and online music distribution, good music is increasingly accessible to many people. The price of good quality tuners, amplifiers, and speakers continues to drop, and driven mostly by demand for home entertainment audio/video systems, audio equipment is increasingly “user-friendly.” The reproduction of theaterquality or live-performance sound in a confined space is challenging enough to be a profession unto itself. It also can be a very rewarding technical hobby for the wellversed electrical engineer. In this Technology Brief, we will cover some of the basics of audio equipment and relate them directly to the concepts taught in this book. Several good audiophile websites exist with more in-depth treatments of these (and other) topics; beyond the audiophile community, the sub-field of audio engineering has an extensive academic and professional literature to consult.

The Basics Reproduced sound starts out as an analog (e.g., the vinyl record) or digital (e.g., the mp3 file) recording. How that

recording is made from real sound with high fidelity is beyond the scope of this Brief (and is a large component of the audio engineering profession). That recording is converted into an electrical signal that is first amplified and then transmitted via cables to speakers. Figure TF24-1 shows a schematic of the process. The audible spectrum of the human ear extends from about 20 Hz to 20 kHz, although the frequency response may vary among different individuals depending on age and other factors. An audio signal is a superposition of many sinusoids oscillating at different frequencies—each with its own individual amplitude. When we say a sound has a lot of bass, for example, we mean that the lowfrequency segment of its spectrum (20 to 100 Hz) has a large amplitude when compared with higher-frequency components. Conversely, very shrill or high-pitched sounds have large-amplitude components in the highfrequency range (10 kHz to 20 kHz). When converting an electrical recording back into the original sound that generated it in the first place, the reproduction fidelity is determined by the degree of distortion that the spectrum undergoes during the playback process. In practice, minimizing spectral distortion can be quite a challenge! Each component of the sound-reproduction system shown in Fig. TF24-1 is characterized by its own transfer function relating its output to its input, and since each of these components is equivalent to a circuit composed of resistive and reactive elements, its transfer function

Figure TF24-1: Schematic of a basic audio-reproduction system.

“book” — 2015/5/4 — 7:20 — page 545 — #46

TECHNOLOGY BRIEF 24: ELECTRICAL ENGINEERING AND THE AUDIOPHILE is bound to exhibit a non-uniform spectral response. The amplifier, for example, may act like a filter, favoring parts of the audible spectrum over others. The cables, which behave (electrically) like the RC transmission line of Fig. 7-35, will favor low-frequency spectral components over high-frequency components. Thus, unless the components of an audio-reproduction system are well designed in order to generate a transfer function with a nearly flat spectral response over the audible range, the reproduced sound will exhibit a distorted spectrum when compared with the original spectrum. While there are many objective metrics by which to judge the fidelity of audio equipment, every listener processes a given sound differently, introducing a subjective component into the experience. A great way to appreciate the concepts introduced in this Technology Brief is to walk into a high-end audiosystems store with three favorite CDs and then to listen to them on many different amplifier-speaker combinations.

Amplifiers It takes quite a bit of power to drive speakers to produce sound in a room.The function of an amplifier is to boost the audio signal’s power high enough to drive the speakers. In doing so, the amplifier must: • keep frequency distortion to a minimum, and • introduce as little noise as possible into the signal. In order to keep frequency distortion to a minimum, the amplifier’s response must be as uniform as possible; in other words, signals of different frequencies and different amplitudes must be amplified with exactly the same gain. To address this, many different transistor–amplifier topologies have been developed over the years. These amplifiers are grouped into classes based on behavior and topology; the principal differences lie in circuit complexity, power consumption, and the degree of fidelity with which the circuit reproduces an input signal. Audio-amplifier circuit topologies are categorized by letter—currently from A to G. Although a description of each class lies beyond the scope of this discussion, very succinct overviews can be found in many places online. In order to reduce the noise during the amplification step, two-amp stages often are used. The first stage is called the pre-amplifier. Preamps have very good noise characteristics and amplify the signal partway (this mid-level signal is called the line signal). Often, this is simply an amplification of the voltage level. The power amp then boosts this signal (which does not have much noise) to a level high enough to drive

545

speakers; this usually requires significant current amplification to provide enough overall power to the speakers.

Cables The cables that transfer a signal between sources, amplifiers, crossovers, and speakers can themselves distort the signal. Cables behave exactly like the transmission line of Fig. 7-35; the distributed resistance and capacitance act like a filter with an associated frequency response. In general, cables should be: • as short as possible, • properly impedance-matched to both the output of the amplifier and the input of the speakers, and • properly terminated so the cable connections to the equipment do not introduce capacitances. All three of these objectives easily are accomplished using industry-standard cables and connectors. In some modern systems, transmission of audio signals between non-speaker components (e.g., from a tuner to an amp or from an amp to a TV) is often performed in digital form so as to eliminate both noise issues and frequency distortion.

Speakers A speaker is any electro-mechanical device or transducer that converts an electrical signal into sound. Electromagnetic transducers are the most commonly used type for consumer audio applications (Fig. TF24-2),

Wire coil

Mechanical suspension (spider)

Magnet

+

iin

υin Zin

_

Cone

Motion

FigureTF24-2: Conceptual illustration of an electromagnetic speaker transducer. The current from the amplifier runs through a coil that induces an electromagnetic force on the cone in proportion to the amplitude of the input signal. The cone motion produces pressure waves and, hence, sound.

“book” — 2015/5/4 — 7:20 — page 546 — #47

546

TECHNOLOGY BRIEF 24: ELECTRICAL ENGINEERING AND THE AUDIOPHILE

Speaker frequency response

Figure TF24-4: Frequency response of a good consumer-quality speaker consisting of a tweeter and woofer.

Figure TF24-3: Electrostatic speakers consist of a very thin (∼ 20 μm) polymer membrane, called a diaphragm, which is coated with a conductor. This membrane is suspended between two perforated electrodes. The diaphragm is held at a dc potential of several kV. The voltage between the electrodes and the diaphragm is driven by the amplified audio signal so as to displace the diaphragm and move air (which produces sound). The principle behind actuation is very similar to that discussed in Technology Brief 10: Micromechanical Sensors and Actuators. Most electrostatic speakers have poor base response, so they are usually paired with subwoofers (like the one shown here). (Image courtesy of MartinLogan, Ltd.)

although several other technologies, such as electrostatic speakers (Fig. TF24-3), exist as well. The principal metric when choosing a speaker is arguably its frequency response (Fig. TF24-4). Ideally, a speaker will provide a very flat response. This means that signals at different frequencies recreated into sound all at the same audio level. Generally speaking, very small speakers have difficulty reproducing very low frequencies (i.e., bass); a deep drum or baseline may be lost entirely when listening through a small speaker.

The most common method for obtaining a nice flat frequency response is to drive several speakers together—each with a different but complementary frequency response. When listened to as a group, the frequency response is close to flat. For example, tweeters are small speakers intended for reproducing high-frequency sound, while woofers only reproduce the lowest frequencies. A common entry-level speaker consists of a tweeter, a mid-range speaker, and a woofer all housed together. With appropriate crossover circuits the ensemble can exhibit a good response.

Crossover circuits As we noted earlier, most speakers cannot handle the entire range of frequencies in the audio range. In order to split the signal for use by the different speakers (such as a tweeter, a mid-range speaker, and a sub-woofer), passive filters are used. The signal is applied to a set of filters that produce three outputs: one output contains only low frequencies in some range, a second contains midrange frequencies and a third output contains only highfrequency harmonics. In this way, each speaker receives a dedicated signal that contains only the frequencies it can reproduce properly. Designing crossovers can be an involved process that takes into account many variables, including the amount of current in the input signal, the input impedances of all of the speakers, and the frequency range of each speaker. Without careful design, the crossover circuit can provide too much signal power to one speaker and too little to another, thereby distorting the overall frequency response heard by the listener.

“book” — 2015/5/4 — 7:20 — page 547 — #48

9-8 APPLICATION NOTE: MODULATION AND THE SUPERHETERODYNE RECEIVER

547

Amplitude modulated waveform

Signal υs(t)

AM

υam(t) = A(t) cos 2πfct, with: A(t) = a0 + υs(t) fc = constant

Carrier υc = A cos 2πfct (a) AM

Frequency modulated waveform

Signal υs(t)

FM

υfm(t) = A cos 2πfct, with: fc = f0 + b0 υs(t) A = constant Carrier υc = A cos 2πfct (b) FM

Figure 9-29: Overview of AM and FM.

9-8 Application Note: Modulation and the Superheterodyne Receiver 9-8.1

Modulation

In the language of electronic communication, the term signal refers to the information to be communicated between two different locations or between two different circuits, and the term

carrier refers to the sinusoidal waveform that carries the information. The latter is of the form υc (t) = A cos 2πfc t,

(9.95)

where A is its amplitude and fc is its carrier frequency. The sinusoid can be used to carry information by modulating (varying) its amplitude—in which case A becomes A(t)—while keeping fc constant. In the example shown in Fig. 9-29(a),

“book” — 2015/5/4 — 7:20 — page 548 — #49

548

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

υam(t) = A(t) cos (2πfct), with: A(t) = a0 + υs(t) Antenna

fc = 1 MHz Tuner RF amp f fc Adjustable filter

fc

Mixer

IF filter/amplifier fIF

+

+

+

υam(t)

υin(t)

υout(t)

_

gRF

_

gRF

_

_ υLO(t)

+

Demodulator

+

+

f υIF(t) fIF

υd(t)

_

gIF

Audio amplifier Speaker

_

gd

fLO

Adjustable local oscillator Tuning knob

RF to IF conversion Figure 9-30: Block diagram of superheterodyne receiver.

multiplication of the signal waveform by the sinusoidal carrier generates an amplitude-modulated (AM) carrier whose envelope is identical to the signal waveform. Alternatively, we can apply frequency modulation (FM) by keeping A constant and varying fc in a fashion that mimics the variation of the signal waveform, as illustrated by Fig. 9-29(b). FM usually requires more bandwidth than AM, but it is also more immune to noise and interference, thereby delivering a higher-quality sound than AM. Many other types of modulation techniques also are available, including phase modulation (changing the phase of the carrier) and pulse-code modulation.

9-8.2 The Superheterodyne Receiver Let us assume the signal υs (t) in Fig. 9-29(a) is an audio signal and the carrier frequency fc = 1 MHz. Let us also assume that the signal was used to generate an amplitude-modulated waveform, which was then fed into a transmit antenna. After propagating through the air along many different directions (as dictated by the antenna radiation pattern), part of the AM waveform was intercepted by a receive antenna connected to an AM receiver. Prior to 1918, the receiver would have been a tuned-radio frequency receiver or a regenerative receiver, both of which suffered from poor frequency selectivity and

low immunity to noise. In either case, the receiver would have demodulated the AM signal by suppressing the carrier and preserving the envelope, thereby retrieving the original signal υs (t) (or more realistically, some distorted version of υs (t)). To overcome the shortcomings of such receivers, Edwin Armstrong introduced the heterodyne receiver in 1918 by proposing the addition of a receiver stage to convert the carrier frequency of the AM signal fc to a fixed lower frequency (now called the intermediate frequency fIF ) before detection (demodulation). [Armstrong also invented frequency modulation in 1935.] The superheterodyne concept proved to be one of the foundational enablers of 20th-century radio transmission. It is still in use in most AM and FM analog receivers, although it slowly is getting supplanted by digital techniques (Section 9-8.4). Figure 9-30 shows a basic block diagram of a superheterodyne receiver. The tuner is a bandpass filter whose center frequency can be adjusted to allow the intended signal at fc = 1 MHz (for example) to pass through, while rejecting signals at other carrier frequencies. After amplification by the radio-frequency (RF) amplifier, the AM signal either can be demodulated directly (which is what receivers did prior to 1918) or it can be converted into an IF signal by mixing (multiplying)

“book” — 2015/5/4 — 7:20 — page 549 — #50

9-8 APPLICATION NOTE: MODULATION AND THE SUPERHETERODYNE RECEIVER it with another locally generated sinusoidal signal provided by a local oscillator. As will be explained in Section 9-8.3, the mixer is a device that multiplies the two signals available at its input and generates an output signal whose frequency is

549

the conversion takes place, consider the general case of two signals given by υin (t) = A(t) cos 2πfc t

(9.97a)

υLO (t) = ALO cos 2πfLO t,

(9.97b)

and fIF = fLO − fc ,

(9.96)

where fLO is the local-oscillator frequency. The frequency conversion given by Eq. (9.96) assumes that fLO ≥ fc ; otherwise, fIF = fc − fLO if fLO < fc . It is important to note that frequency conversion changes the carrier frequency of the AM waveform from fc to fIF , but the audio signal υs (t) remains unchanged; it merely is getting carried by a different carrier frequency. The diagram in Fig. 9-30 indicates that the tuning knob controls the center of the adjustable tuner as well as the local oscillator frequency. By synchronizing these two frequencies to each other, the IF frequency remains always a constant. This is an important feature of the superheterodyne receiver, because it insures that the same IF filter/amplifier can be used to provide high-selectivity filtering and high-gain amplification, regardless of the carrier frequency of the AM signal. In the AM radio band, the carrier frequency of the audio signals transmitted by an AM radio station may be at any frequency between 530 and 1610 kHz. Because of the built-in synchronization between the tuner and the local oscillator, the IF frequency of an AM receiver is always at 455 kHz, which is the standard IF for AM radio. Similarly, the standard IF for FM radio is 10 MHz, and the standard IF for television is 45 MHz. It is impractical to design and manufacture high-performance components at every frequency in the radio spectrum. By designating certain frequencies as IF standards, industry was able to develop devices and systems that operate with very high performance at those frequencies. Consequently, frequency conversion to an IF band is very prevalent not only in radio and TV receivers but also in radar sensors, satellite communication systems and transponders, among others.

where A(t) represents the amplitude of the audio signal waveform, υs (t) (Fig. 9-29(a)), and ALO is a constant amplitude associated with the local oscillator (LO) signal. A mixer is a diode circuit that has two inputs and one output with its output voltage υout (t) being equal to the product of its input voltages: υout (t) = υin (t) × υLO (t) = A(t) ALO cos 2πfc t cos 2πfLO t. (9.98) Application of the trigonometric identity cos x cos y = 21 [cos(x + y) + cos(x − y)]

(9.99)

leads to υout (t) =

A(t) ALO cos[2π(fc + fLO )t] 2 A(t) ALO + cos[2π(fLO − fc )t]. 2

(9.100)

Let us consider the case where fc = 1 MHz and fLO = 1.445 MHz. The expression for υout (t) becomes υout (t) = A (t) cos 2πfs t + A (t) cos 2πfd t,

(9.101)

where A(t) ALO , (9.102) 2 and fs and fd are the sum and difference frequencies: A (t) =

fs = fc + fLO = 2.445 MHz

(9.103a)

fd = fLO − fc = 0.445 MHz.

(9.103b)

and

9-8.3

Frequency Conversion

Regardless of the specific type of modulation used in a modern communication system, it will employ one or more steps of frequency conversion, whereby the carrier frequency is changed from an initial frequency f1 to a new frequency f2 . If f2 is higher than f1 , it is called up-conversion, and the reverse is called down-conversion. In the AM example of Fig. 9-30, f1 = fc = 1 MHz and f2 = fIF = 455 kHz. To explain how

Thus, υout (t) consists of two signal components with markedly different carrier frequencies. By selecting a narrow IF filter/ amplifier in Fig. 9-30 with a center frequency fIF = fd , only the difference-frequency component of υout (t) will make it through the filter. Consequently, its output is given by υIF (t) = gIF A (t) cos 2πfIF t,

“book” — 2015/5/4 — 7:20 — page 550 — #51

550

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

where gIF is the voltage gain factor of the IF filter/amplifier. Demodulation, which is a low-frequency filtering process, removes the IF carrier, leaving behind a detected signal given by υd (t) = gd gIF A (t), where gd is a demodulator constant. Since A (t) is directly proportional to the original audio signal υs (t), υd (t) becomes (ideally) a replica of υs (t). Figure 9-31: A series RLC filter implemented in Multisim.

9-8.4

Software Radio

The steady increase in the speed of digital circuits has made it possible to perform all of the functions of a superheterodyne receiver directly in the digital domain. Some implementations consist of little more than an antenna connected to the input pin of a digital chip. The chip converts the input signal into digital format and then performs all of the mixing, filtering, amplifying, and demodulating functions by direct computation. This digital approach, first proposed in the 1980s, is sometimes called software radio. In practice, analog-to-digital converters do not usually have the specifications required to directly sample signals coming from an antenna, and low-noise amplifiers are needed at the front end of the digital system. Additionally, many software radio implementations still use mixers at the front end, and simply digitize the signal coming out of the mixer. Concept Question 9-18: What are the advantages of FM

over AM? (See

)

a convenient way to evaluate the frequency response of a circuit using Multisim. These tools are illustrated in the next three examples. Example 9-13: RLC Circuit

Design a series RLC bandpass filter with a center frequency of 10 MHz and Q = 50. Use Multisim to generate magnitude and phase plots covering the range from 8 to 12 MHz. Solution: The specified filter can be designed with an infinite number of different combinations of R, L, and C. We will choose a realistic value for L, namely 0.1 mH, which will dictate that C be C= =

Concept Question 9-19: What is the fundamental

contribution of the superheterodyne receiver, and why is it significant? (See )

1 ω02 L 1 (2π

)

× 10−4

= 2.53 pF. Next, we select the value of R to satisfy the requirement on Q. From Table 9-3, we obtain

Concept Question 9-20: What does a mixer do?

(See

× 107 )2

R=

ω0 L Q

2π × 107 × 10−4 50 = 125.7 .

=

9-9 Spectral Response with Multisim The AC Analysis and Parameter Sweep tools are very useful when analyzing the frequency response of a circuit. The Network Analyzer, first introduced in Chapter 8, also provides

With all three elements specified, we construct the Multisim circuit shown in Fig. 9-31. Before performing AC Analysis, we should double-click on the ac source and change its value to 1 V

“book” — 2015/5/4 — 7:20 — page 551 — #52

9-9

SPECTRAL RESPONSE WITH MULTISIM

551

Figure 9-32: Magnitude and phase plots for the circuit of Fig. 9-31 generated by AC Analysis for Example 9-13.

(rms). Next, we select Simulate → Analyses → AC Analysis, and then we set FSTART to 8 MHz and FSTOP to 12 MHz. With both the Sweep Type and Vertical Scale set to Linear, the number of points set to 1000, and the variable selected for analysis is V(3), AC Analysis generates the plots displayed in Fig. 9-32. The magnitude plot exhibits a peak at 10 MHz, and the phase goes through 0◦ at that frequency. To verify that the circuit has a Q = 50, we use the cursors to establish √ the locations at which the vertical value of the curve is 1/ 2 = 0.707 V. The separation between the two cursors (labeled “dx” in the cursor box) is 200.0699 kHz. This is the half-power bandwidth B. The quality factor is Q= =

Example 9-14: Parameter Sweep

Apply Parameter Sweep to the circuit in Fig. 9-31 to generate spectral responses for C1 = 1 pF, 4 pF, 7 pF, and 10 pF.

ω0 B

Solution: Starting with the circuit in Fig. 9-31, we set V1 = 1 V. Next, we select Simulate → Analyses → Parameter Sweep. Upon selecting the parameter we wish to vary (capacitance C1), its minimum (1 pF), maximum (10 pF), step size (3 pF), and number of points (4), we select AC Analysis in the More Options box. This allows us to set the frequency range, the type of sweep (linear), and the number of points—just as we did previously in Example 9-13, except that the frequency range is 0 to 20 MHz. The Simulate command generates the plots shown in Fig. 9-33.

107 200.0699 × 103

Example 9-15: Bode Plots

= 49.98, which is approximately equal to the specified value.

Reproduce the circuit of Fig. 9-26(a) in Multisim and generate Bode plots corresponding to the outputs of the three stages.

“book” — 2015/5/4 — 7:20 — page 552 — #53

552

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

C1 = 7 pF C1 = 10 pF

C1 = 4 pF C1 = 1 pF

Figure 9-33: AC analysis plots for the circuit in Fig. 9-31 generated with the Parameter Sweep tool in Example 9-14. The capacitance was varied from 1 to 10 pF.

Solution: The circuit is reproduced in Fig. 9-34(a), and part (b) displays the results. In order to generate these plots, we use AC Analysis with FSTART = 104 (rad/s)/2π = 1.592 kHz and FSTOP = 107 /2π = 1.592 MHz. The number of points was set to 200, Sweep Type = Decade, and Vertical Scale = Decibel. Example 9-16: Bode Plotter Instrument

Use the Bode Plotter Instrument to generate magnitude and phase plots of the circuit in Fig. 9-31 over the frequency range of 8 to 12 MHz.

Solution: Go to Simulate → Instruments → Bode Plotter. Connect the “IN” terminals across the V1 source and connect the “OUT” terminals across the resistor R1. Bring up the Bode Plotter Instrument window. With the Magnitude Mode selected, set the horizontal scale to Lin (for linear), set I (initial frequency) to 8 MHz, and set F (final frequency) to 12 MHz. For the vertical scale, leave it on Log, set I to −50 dB, and set F to 5 dB. Select the Phase mode and for the vertical scale set I to −100 deg and F to 100 deg. Run the Interactive Simulation by pressing F5 or the appropriate button or toggle switch on the toolbar. In the Magnitude and Phase mode, you will generate plots similar to those shown in Fig. 9-35(a) and (b), respectively.

“book” — 2015/5/4 — 7:20 — page 553 — #54

SPECTRAL RESPONSE WITH MULTISIM

553

1 mVrms

(a) Three-stage circuit of Fig. 9-26(a) 15 −10 (dB)

9-9

−35 Single stage −60 Two stages

−85 −110

Three stages

−135 −160

(b) Bode plots Figure 9-34: Three-stage op-amp circuit of Fig. 9-26(a) reproduced in Multisim for Example 9-15.

“book” — 2015/5/4 — 7:20 — page 554 — #55

554

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

(a) Magnitude plot

(b) Phase plot Figure 9-35: Output of Bode Plotter Instrument for Example 9-16.

“book” — 2015/5/4 — 7:20 — page 555 — #56

9-9

SPECTRAL RESPONSE WITH MULTISIM

555

Summary Concepts • The transfer function of a circuit is the ratio of a phasor output voltage or current to a phasor input voltage or current. • The transfer function is characterized by magnitude and phase plots describing the spectral response of the circuit. • At the resonant frequency ω0 , the input impedance of the circuit is purely real.

• The order of a filter defines the gain roll-off rate of the magnitude spectrum in the stopband. • Active filters are used primarily at frequencies below 1 MHz, whereas passive filters are better suited at higher frequencies. • Active filters can provide power gain, and they easily can be cascaded in series or in parallel to generate the desired frequency response.

• The Bode diagram uses straight-line approximations on a semilog-scale to display the magnitude and phase spectra of the transfer function. • The quality factor Q of a bandpass filter defines the degree of frequency selectivity of the filter.

• In a superheterodyne receiver, the RF frequency is converted into an IF frequency for amplification and filtering prior to demodulation. • Parameter sweep can be used in Multisim to compare the circuit response for different values of a key parameter.

Mathematical and Physical Models Resonant Frequency ω0 {Zin (ω)} = 0

Series and Parallel Bandpass RLC Filters

@ ω = ω0

ω0 =

Magnitude and Frequency Scaling R  = Km R, C =

1 C, Km Kf

L =

Km L Kf

ω  = Kf ω

ωc1 ωc2 = √

ω0 L Q= R R Q= ω0 L

1 LC

(series) (parallel)

Active Filters

dB Scale If G = XY If G =



X Y

Important Terms −3 dB frequency active filter amplitude modulation AM radio band AND logic gate bandpass filter

G [dB] = X [dB] + Y [dB]

Sections 9-6 and 9-7

G [dB] = X [dB] − Y [dB]

Provide definitions or explain the meaning of the following terms: bandreject filter bandwidth Bode plot Bode diagram carrier carrier frequency

connected in parallel connected in series corner frequency cutoff frequency damping factor dc gain

degree of selectivity demodulate down-conversion Edwin Armstrong energy dissipated exact plots

“book” — 2015/5/4 — 7:20 — page 556 — #57

556

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

Important Terms (continued) factor filter order frequency conversion frequency modulation frequency response frequency scaling frequency scaling factor frequency-selective circuits gain factor gain roll-off rate half-power frequency heterodyne receiver high-frequency gain highpass filter idealized response IF IF filter/amplifier

intermediate frequency local oscillator lowpass filter magnitude magnitude scaling magnitude scaling factor magnitude response mix mixer normalized power order OR gate passband passive filter phase angle phase response pole

PROBLEMS

pole @ origin factor pole factor practical circuit prototype model quadratic-pole factor quadratic-zero factor quality factor radio-frequency regenerative receiver relative power resonance condition resonant frequency RF signal simple-pole factor simple-zero factor software radio

standard form store straight-line approximation stopband transfer function trivial resonance tuned-radio frequency receiver tuner unity input up-conversion voltage transfer function zero zero factor zero @ origin factor

*9.3 Determine the resonant frequency of the circuit shown in Fig. P9.3, given that R = 1 k, L = 10 mH, and C = 10 nF.

Section 9-1: Transfer Function *9.1 Determine the resonant frequency of the circuit shown in Fig. P9.1, given that R = 100 , L = 5 mH, and C = 1 μF.

L

R

L

C

Figure P9.3: Circuit for Problem 9.3.

C

R

Figure P9.1: Circuit for Problem 9.1.

9.4 Determine the resonant frequency of the circuit shown in Fig. P9.4, given that R = 1 k, L = 10 mH, and C = 10 nF.

9.2 Determine the resonant frequency of the circuit shown in Fig. P9.2, given that R = 100 , L = 5 mH, and C = 1 μF.

L R

C

C L

R

Figure P9.2: Circuit for Problem 9.2.



Answer(s) available in Appendix G.

Figure P9.4: Circuit for Problem 9.4.

*9.5 Determine the resonant frequency of the circuit shown in Fig. P9.5, given that R1 = 10 , R2 = 100 , L = 5 mH, and C = 0.1 μF.

“book” — 2015/5/4 — 7:20 — page 557 — #58

PROBLEMS

557

R2



L

R1

C

Vs

1F

0.3 F

0.04 H

0.04 H

1F

+

+ _



Vo

_

Figure P9.5: Circuit for Problem 9.5. Figure P9.9: Circuit for Problem 9.9.

9.6 For the circuit shown in Fig. P9.6, determine (a) the transfer function H = Vo /Vi and (b) the frequency ωo at which H is purely real.

C

_

(a) Obtain an expression for the input impedance Zin (ω).

(c) Scale the circuit by Km = 20 and write down the new expression for the input impedance.

+

L1

Vi

R

Vo

_

(d) Is the value of ω at which the input impedance of the scaled circuit is real the same or different from the answer of part (b)?

Figure P9.6: Circuit for Problem 9.6.

Zin

R1 + _

C

R1

9.7 For the circuit shown in Fig. P9.7, determine (a) the transfer function H = Vo /Vi and (b) the frequency ωo at which H is purely real.

Vi

For the circuit shown in Fig. P9.10:

(b) If R1 = R2 = 1 , C = 1 F, and L = 5 H, at what angular frequency is Zin purely real?

L2

+

9.10

R2

L

Figure P9.10: Circuit for Problem 9.10.

L

C

R2

+ Vo

_

9.11

For the circuit shown in Fig. P9.11:

(a) Obtain an expression for the input impedance Zin (ω). Figure P9.7: Circuit for Problem 9.7.

*(b) If R1 = 1 , R2 = R3 = 2 , L = 1 H, and C = 1 F, at what angular frequency is Zin purely real? (c) Redraw the circuit after scaling it by Km = 103 and Kf = 105 . Specify the new element values.

Section 9-2: Scaling *9.8 What values of the scaling factors Km and Kf should be applied to scale a circuit containing a 1 F capacitor and 4 H inductor into one containing 1 μF and 10 mH, respectively? 9.9 The corner frequency of the highpass-filter circuit shown in Fig. P9.9 is approximately 1 Hz. Scale the circuit up in frequency by a factor of 105 while keeping the values of the inductors unchanged.

C

R1 L

R3

Zin R2

Figure P9.11: Circuit for Problem 9.11.

“book” — 2015/5/4 — 7:20 — page 558 — #59

558

CHAPTER 9

9.12 Circuit (b) in Fig. P9.12 is a scaled version of circuit (a). The scaling process may have involved magnitude or frequency scaling, or both simultaneously. If R1 = 1 k gets scaled to R1 = 10 k, supply the impedance values of the other elements in the scaled circuit.

j10 Ω L1

j20 Ω −j5 Ω L2

C1 R1

C2 R2

1 kΩ

−j50 Ω 2 kΩ

(a) Original circuit

L1

L2

C1 R1

10 kΩ

C2 R2

(b) Scaled circuit Figure P9.12: Circuits for Problem 9.12.

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS (b) 0.4 dB *(c) −12 dB (d) −66 dB 9.17 Generate Bode magnitude and phase plots (straight-line approximations) for the following voltage transfer functions. j 100ω (a) H(ω) = 10 + j ω 0.4(50 + j ω)2 (j ω)2 (40 + j 80ω) (c) H(ω) = (10 + j 50ω) (20 + j 5ω)(20 + j ω) (d) H(ω) = jω (b) H(ω) =

9.18 Generate Bode magnitude and phase plots (straight-line approximations) for the following voltage transfer functions. 30(10 + j ω) (a) H(ω) = (200 + j 2ω)(1000 + j 2ω) j 100ω (b) H(ω) = (100 + j 5ω)(100 + j ω)2 (200 + j 2ω) (c) H(ω) = (50 + j 5ω)(1000 + j ω) 9.19 Generate Bode magnitude and phase plots (straight-line approximations) for the following voltage transfer functions.

Section 9-3: Bode Plots 9.13 Convert the following power ratios to dB. (a) 3 × 102 *(b) 0.5 × 10−2 √ (c) 2000 (d) (360)1/4 9.14 Convert the following power ratios to dB. *(a) 6e3 (b) 2.3 × 103 + 60 (c) 24(3 × 107 ) (d) 4/(5 × 103 ) 9.15 Convert the following voltage ratios to dB. (a) 2 × 10−4 (b) 3000 √ (c) 30 (d) 6/(5 × 104 ) 9.16 Convert the following dB values to voltage ratios. (a) 46 dB

(a) H(ω) =

4 × 104 (60 + j 6ω) (4 + j 2ω)(100 + j 2ω)(400 + j 4ω)

(b) H(ω) =

(1 + j 0.2ω)2 (100 + j 2ω)2 (j ω)3 (500 + j ω)

(c) H(ω) =

8 × 10−2 (10 + j 10ω) j ω(16 − ω2 + j 4ω)

(d) H(ω) =

4 × 104 ω2 (100 − ω2 + j 50ω) (5 + j 5ω)(200 + j 2ω)3

9.20 Generate Bode magnitude and phase plots (straight-line approximations) for the following voltage transfer functions. j 5 × 103 ω(20 + j 2ω) (2500 − ω2 + j 20ω) 512(1 + j ω)(4 + j 40ω) (b) H(ω) = (256 − ω2 + j 32ω)2 (a) H(ω) =

(c) H(ω) =

j (10 + j ω) × 108 (20 + j ω)2 (500 + j ω)(1000 + j ω)

*9.21 Determine the voltage transfer function H(ω) corresponding to the Bode magnitude plot shown in Fig. P9.21. The phase of H(ω) approaches 180◦ as ω approaches 0.

“book” — 2015/5/4 — 7:20 — page 559 — #60

PROBLEMS

559 9.24 Determine the voltage transfer function H(ω) corresponding to the Bode magnitude plot shown in Fig. P9.24. The phase of H(ω) is −90◦ at ω = 0.

M [dB] 26 dB 20 dB

M [dB] 36 dB 30 dB

10 dB 6 dB 0

0.1

1

100

1000

ω (rad/s)

20 dB

Figure P9.21: Bode magnitude diagram for Problem 9.21.

0 9.22 Determine the voltage transfer function H(ω) corresponding to the Bode magnitude plot shown in Fig. P9.22. The phase of H(ω) is 90◦ at ω = 0.

100 200

2000 4000

ω (rad/s)

Figure P9.24: Bode magnitude plot for Problem 9.24.

9.25 Determine the voltage transfer function H(ω) corresponding to the Bode magnitude plot shown in Fig. P9.25. The phase of H(ω) is 0◦ at ω = 0.

M [dB] 60 dB

M [dB] 40 dB 20 dB 14 dB

20 dB 0

0.5

5

50

500

ω (rad/s)

0 −6 dB

5

50

500

ω (rad/s)

Figure P9.22: Bode magnitude plot for Problem 9.22.

−20 dB *9.23 Determine the voltage transfer function H(ω) corresponding to the Bode magnitude plot shown in Fig. P9.23. The phase of H(ω) is 180◦ at ω = 0.

−40 dB −46 dB

M [dB] Figure P9.25: Bode magnitude plot for Problem 9.25.

20 dB −20 dB/decade

Sections 9-4 and 9-5: Passive Filters

10 dB 0

20 dB/decade

2

10 20

100

ω (rad/s)

Figure P9.23: Bode magnitude plot for Problem 9.23.

9.26 The element values of a series RLC bandpass filter are R = 5 , L = 20 mH, and C = 0.5 μF. *(a) Determine ω0 , Q, B, ωc1 , and ωc2 . (b) Is it possible to double the magnitude of Q by changing the values of L and/or C while keeping ω0 and R unchanged? If yes, propose such values, and if no, why not?

“book” — 2015/5/4 — 7:20 — page 560 — #61

560

CHAPTER 9

9.27 A series RLC bandpass filter has half-power frequencies at 1 kHz and 10 kHz. If the input impedance at resonance is 6 , what are the values of R, L, and C? 9.28 A series RLC circuit is driven by an ac source with a phasor voltage Vs = 10 30◦ V. If the circuit resonates at 103 rad/s and the average power absorbed by the resistor at resonance is 2.5 W, determine the values of R, L, and C, given that Q = 5. *9.29 The element values of a parallel RLC circuit are R = 100 , = L = 10 mH, and C = 0.4 mF. Determine ω0 , Q, B, ωc1 , and ωc2 . 9.30 Design a parallel RLC filter with f0 = 4 kHz, Q = 100, and an input impedance of 25 k at resonance. 9.31 For the circuit shown in Fig. P9.31: (a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = 1 , R2 = 2 , C1 = 1 μF, and C2 = 2 μF. (c) Determine the cutoff frequency ωc and the slope of the magnitude (in dB) when ω/ωc  1 and when ω/ωc  1.

C1

Vi

_

9.33 For the circuit shown in Fig. P9.33: (a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Generate spectral plots for the magnitude and phase of H(ω), given that R = 100 , L = 0.1 mH, and C = 1 μF. (c) Determine the cutoff frequency ωc and the slope of the magnitude (in dB) when ω/ωc  1.

L

+

+ C

Vi

_

R

Vo

_

Figure P9.33: Circuit for Problem 9.33.

9.34 For the circuit shown in Fig. P9.34: (a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Generate spectral plots for the magnitude and phase of H(ω), given that R = 10 , L = 1 mH, and C = 10 μF. (c) Determine the cutoff frequency ωc and the slope of the magnitude (in dB) when ω/ωc  1.

C2

R1

+

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

R2

+ Vo

_

C

+

+ L

Vi

_

R

Vo

_

Figure P9.31: Circuit for Problem 9.31. Figure P9.34: Circuit for Problem 9.34.

9.32 For the circuit shown in Fig. P9.32: (a) Obtain an expression for H(ω) = Vo /Vi in standard form. *(b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = 1 , R2 = 2 , L1 = 1 mH, and L2 = 2 mH. (c) Determine the cutoff frequency ωc and the slope of the magnitude (in dB) when ω/ωc  1 and when ω/ωc  1.

L2

R1

+ Vi

_

9.35 For the circuit shown in Fig. P9.35: *(a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Generate spectral plots for the magnitude and phase of H(ω), given that R = 50  and L = 2 mH. (c) Determine the cutoff frequency ωc and the slope of the magnitude (in dB) when ω/ωc  1.

L1

R R2

+

+

Vo

Vi

Figure P9.32: Circuit for Problem 9.32.

_

_

+ L

R

Vo

Figure P9.35: Circuit for Problem 9.35.

_

“book” — 2015/5/4 — 7:20 — page 561 — #62

PROBLEMS 9.36

561

For the circuit shown in Fig. P9.36:

C

(a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Generate spectral plots for the magnitude and phase of H(ω), given that R = 50  and L = 2 mH.

R1

_ R

+ Vi

_

R2

+ R

L

Vs

+ _

+

+ Vo

_

Vo

_

Figure P9.38: Circuit for Problem 9.38. Figure P9.36: Circuit for Problem 9.36.

9.39

*(a) Obtain an expression for H(ω) = Vo /Vi in standard form.

Sections 9-6 and 9-7: Active Filters 9.37

For the op-amp circuit of Fig. P9.39:

(b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = R2 = 100 , C1 = 10 μF, and C2 = 0.4 μF.

For the op-amp circuit of Fig. P9.37:

*(a) Obtain an expression for H(ω) = Vo /Vs in standard form.

(c) What type of filter is it? What is its maximum gain?

(b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = 1 k, R2 = 4 k, and C = 1 μF. (c) What type of filter is it? What is its maximum gain?

R1 R2 Vs

_ R1 C

Vs

+ _

+

+

_

R2

+ C2

+ Vo

_

Figure P9.39: Circuit for Problems 9.39 and 9.40.

Vo

_

Figure P9.37: Circuit for Problem 9.37.

9.38

+ _

C1

For the op-amp circuit of Fig. P9.38:

9.40 Repeat Problem 9.39 after interchanging the values of C1 and C2 to C1 = 0.4 μF and C2 = 10 μF. 9.41

For the op-amp circuit of Fig. P9.41:

(a) Obtain an expression for H(ω) = Vo /Vs in standard form.

*(a) Obtain an expression for H(ω) = Vo /Vs in standard form.

(b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = 99 k, R2 = 1 k, and C = 0.1 μF.

(b) Generate spectral plots for the magnitude and phase of H(ω), given that R1 = 1 k, R2 = 20 , C1 = 5 μF, and C2 = 25 nF.

(c) What type of filter is it? What is its maximum gain?

(c) What type of filter is it? What is its maximum gain?

“book” — 2015/5/4 — 7:20 — page 562 — #63

562

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

C2

R2 R1

R2 R1 Vs

C1

_ +

+ _

+

Vs

+ _

_ +

+

V0

C

Vo

_

_

Figure P9.46: Circuit for Problem 9.46. Figure P9.41: Circuit for Problem 9.41.

9.42 Design an active lowpass filter with a gain of 4, a corner frequency of 1 kHz, and a gain roll-off rate of −60 dB/decade. 9.43 Design an active highpass filter with a gain of 10, a corner frequency of 2 kHz, and a gain roll-off rate of 40 dB/decade. 9.44 Show that the transfer function of the circuit shown in Fig. P9.44 is given by   ω Vo , = −G 1 + j H(ω) = Vs ωc

Vs

9.49 Using the circuit layout shown in Fig. 9-15, design a tuner that uses a variable inductor, a capacitor, and a resistor. The input impedance of the tuner should be 377  at 1 MHz, and its bandwidth should be 2 percent. *9.50 What range of frequencies should the local oscillator be able to provide to mix the FM radio range (88 to 108 MHz) down to 10 MHz?

R2

_ R1

+

9.48 The element values in the circuit of the second-order bandpass filter shown in Fig. P9.48 are Rf1 = 100 k, Rs1 = 10 k, Rf2 = 100 k, Rs2 = 10 k, Cf1 = 3.98 × 10−11 F, and Cs2 = 7.96 × 10−10 F. Generate a spectral plot for the magnitude of H(ω) = Vo/Vs. Determine the frequency locations of the maximum value of M [dB] and its half-power points.

Section 9-8: Superheterodyne Receiver

and relate G and ωc to R1 , R2 , and C.

C

9.47 Use resistors, capacitors, and a single op amp to design a circuit with input voltage Vs , output voltage Vo , and transfer function  Vo ω  H(ω) = = −10 1 + j . Vs 100

V0

Figure P9.44: Circuit for Problem 9.44.

9.45 Repeat Problem 9.41 after replacing the series combination of R1 and C1 with a parallel combination. *9.46 Consider the circuit shown in Fig. P9.46. Obtain its transfer function H(ω) = Vo /Vs for R1 = 1 k, R2 = 10 k, and C = 1 μF. What role, if any, does the capacitor play? Explain.

Section 9-9: Multisim

9.51 Generate plots in Multisim for the magnitude and phase of the transfer function for a series bandpass filter with L = 1 mH, f0 = 1 MHz, and Q = 10. Choose FSTART = 100 kHz and FSTOP = 10 MHz. 9.52 Perform a Parameter Sweep in Multisim for capacitor Cs2 of the two-stage bandpass filter shown in Fig. 9-27. Generate response plots from 10 Hz to 100 kHz for each of five equally spaced values of Cs2 starting at 1 nF and ending at 15 nF. 9.53 Use Multisim to generate spectral plots for the magnitudes and phases of voltages VC and Vo in the circuit of

“book” — 2015/5/4 — 7:20 — page 563 — #64

PROBLEMS

563

Cf1

Cf1

Rf1 Rs1

Vs

Rf1

_

Rs1

_

+

+ _

Rs2

Cs2

+

Rf2

_

Rs2

Rf2

Cs 2

_

+

+

+

Vo

_

Figure P9.48: Circuit for Problem 9.48. Fig. P9.53. The circuit is a second-order passive lowpass filter followed by an active highpass filter. Use the following element values: R1 = 20.3 , R2 = R3 = 1.592 k, L1 = 100 nH, C1 = C2 = 1 nF, and υs (t) = cos 2πf t V.

R1

L1

+ Vs _

R2

VC

3 8

R3

C2

+

+ _

9.54 For the circuit in Fig. P9.54, use Multisim to generate spectral plots for the magnitude and phase of H(ω) = Vo /Vi over the range from 100 Hz to 100 kHz. When performing the AC Analysis, use 200 points per decade. Determine the frequencies at which M [dB] is a maximum or a minimum.

+ Vi

_

10−2 H

10−2 H

1 kΩ

1 2

+ 1 2

μF

μF

Vo

_

9.55 For the circuit in Fig. P9.55, use Multisim to generate spectral plots for the magnitude and phase of H(ω) = Vo /Vi

_

9.56 For the circuit in Fig. P9.56, use Multisim to generate spectral plots for the magnitude and phase of H(ω) = Vo /Vi over the range from 100 Hz to 10 kHz. When performing the AC Analysis, use 200 points per decade. Determine the frequencies at which M [dB] is a maximum or a minimum.

8 μF

8 μF

+ Vi

Figure P9.54: Circuit for Problem 9.54.

1 kΩ Vo

Figure P9.55: Circuit for Problem 9.55.

+

0.5 μF 1 8

Vi

10−2 H

_

Figure P9.53: Circuit for Problem 9.53.

1 12

1 6

+

Vo

10−2 H

μF

_

C1

1 12

over the range from 1 to 15 kHz. When performing the AC Analysis, use 104 points in linear scan. Determine the frequencies at which M [dB] is a maximum or a minimum.

_

+ 0.15 H 16 15

100 Ω

μF

Figure P9.56: Circuit for Problem 9.56.

Vo

_

“book” — 2015/5/4 — 7:20 — page 564 — #65

564

CHAPTER 9

FREQUENCY RESPONSE OF CIRCUITS AND FILTERS

9.57 Figure P9.57 depicts a band-stop filter composed of a high-pass filter, a low-pass filter, and a summing amplifier. Construct it in Multisim using the values R = 5 k, CLP = 26 nF, and CHP = 1 nF. Using Multisim’s AC Analysis, plot the transfer function from 100 Hz to 100 kHz for the highpass component alone, the low-pass component alone, and the overall filter all on the same graph. Find the frequency where minimum gain occurs for the overall filter.

R

+ υs

L

~+_

υout C

_

CLP Figure m9.1 Circuit for Problem m9.1.

R R

_

R

+

R

Vin

_ CHP

R R

_

component values are R = 1 , C = 0.5505 F.

Vout

+ R

+ Figure P9.57: Circuit for Problem 9.57.

9.58 Build the circuit shown in Fig. 9-15 in Multisim with values C = 1 pF and R = 377 . Simulate the circuit with L = 5 mH, 10 mH, and 15 mH. Plot the output of the filter at the three tunings on the same plot from 100 kHz to 100 MHz. Potpourri Questions

L = 1.817 H, and

(a) Apply magnitude and frequency scaling to the bandreject filter so that R  = 100  and L = 33 mH. Draw the finished circuit diagram. (b) Determine the center frequency in Hz of the scaled bandreject filter. m9.2

Bode Plots: For the circuit in Fig. m9.2:

(a) Determine the voltage transfer function H(ω) of the filter circuit. Write your finished result in standard form for creating a Bode plot. (b) Substitute ω = 2πf to express the voltage transfer function in terms of oscillation frequency f in Hz. (c) Generate Bode magnitude and phase plots for H(f ) using oscillation frequency f as the independent variable. Use the following component values: R1 = 3.3 k, R2 = 10 k, C1 = 0.01 μF, and C2 = 0.1 μF.

9.59 What role does phase play in the operation of noisecancelling headphones?

R2

9.60 What does high-frequency filtering do to an image? What about low-frequency filtering?

C2 R1

9.61 What is the range of the audible spectrum? Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m9.1 Scaling: Figure m9.1 shows a prototype bandreject filter with center frequency ω0 = 1 rad/s. The prototype

C1 υs

_ +

~+_ Figure m9.2 Circuit for Problem m9.2.

υout

“book” — 2015/5/4 — 7:20 — page 565 — #66

PROBLEMS

565

(d) Determine the following filter circuit properties by inspecting the Bode plot: (1) Low-frequency asymptotes for magnitude and phase (2) High-frequency asymptotes for magnitude and phase (3) Corner frequencies (this filter circuit has two such frequencies) m9.3 Filter Order: The filter circuit shown in Fig. m9.3 uses the component values R = 1.0 k and C = 1.0 μF. (a) Obtain an expression for H(ω) = Vo /Vi in standard form. (b) Substitute ω = 2πf to express H(ω) in terms of the oscillation frequency f in Hz. (c) Generate spectral plots for the magnitude and phase of H(f ). (d) Determine the cutoff frequency fc .

(1) Low-frequency passband corner in Hz, (2) High-frequency passband corner in Hz, and (3) Passband gain in dB. m9.5 Bode Plot for an RLC Circuit I (a) Determine the transfer function of the RLC circuit in Fig. m9.5. Compute the magnitude and phase of the transfer function at 100, 1,000, 5,000, and 10,000 Hz. (b) Using Multisim and myDAQ, capture the Bode plot for the circuit. How does your answer from part (b) verify your answer from part (a)? (c) Determine υout when the input is 0.5 cos(200π t) V. (d) Determine υout when the input is 0.5 cos(103 π t) V.

+ υin(t)

+

_

L

1.5 Ω

3.3 mH

+ C

1 μF

_

C

υi

R

+ R

υout

_

Figure m9.5 Circuit for Problem m9.5.

υo

_

Figure m9.3 Circuit for Problem m9.3.

m9.4 Cascaded Active Filters: A telephone line provides sufficient bandwidth (3 kHz) for intelligible voice conversations, but human hearing has a much higher bandwidth, typically 20 Hz to 20,000 Hz

(a) Design an active bandpass filter to mimic the bandwidth of a telephone line subject to the following constraints: (1) Cascade a first-order active lowpass filter and a first-order active highpass filter,

m9.6 Bode Plot for an RLC Circuit II The circuit in Fig. m9.6 is an RLC circuit. For this problem, there are separate directions for the handwritten, Multisim, and myDAQ portions. (a) The transfer function for the circuit in Fig. m9.6 is given by j ωRC H(ω) = . (1 − ω2 LC) + j ωRC Using the transfer function, compute the magnitude and phase at each of the following frequencies: 100, 500, 1,000, 2,000, 5,000, and 10,000 Hz. Plot the magnitude and phase on separate graphs. (b) For the Multisim and myDAQ portions of this problem, capture a Bode plot for this circuit. Do your answers from part (a) and part (b) agree? (c) Determine υout when υin = 1 cos(5000π t) V.

(2) Set the corner frequencies to 300 Hz and 3.0 kHz, (3) Set the passband gain to 0 dB,

L

(4) Choose resistors in the range 1.0 k to 100 k, and

+

(5) Use the total of four fixed-value resistors and two fixed-value capacitors selected from the parts listed in Appendix A of the tutorial on .

υin(t)

Draw the schematic diagram of your finished design. (b) Predict the performance of your finished design by calculating the following values:

C

+ 33 mH

10 μF

R

1 kΩ

υout

_

_ Figure m9.6 Circuit for Problem m9.6.

“book” — 2015/5/4 — 7:23 — page 566 — #1

10

CHAPTER

Three-Phase Circuits Contents 10-1 10-2 10-3 10-4 TB25 10-5 TB26 10-6 10-7

Overview, 567 Balanced Three-Phase Generators, 568 Source-Load Configurations, 572 Y-Y Configuration, 574 Balanced Networks, 576 Miniaturized Energy Harvesting, 577 Power in Balanced Three-Phase Networks, 582 Inside a Power Generation Station, 586 Power-Factor Compensation, 588 Power Measurement in Three-Phase Circuits, 591 Summary, 595 Problems, 596

Objectives

All voltage values are rms

Large factory Transmission lines 48,000 V 3-phase

480,000 V 3-phase

Primary substation

Nuclear power station

Transmission lines 48,000 V 3-phase

20:1 step-down pole transformer

Distribution lines 4,800 V 3-phase

Learn to: 





Service wire 120 V / 240 V single phase

Analyze both balanced and unbalanced threephase circuits. Convert a Y-source configuration into a -source configuration, and vice versa. Convert a Y-load configuration into a -load configuration, and vice versa.



Compute complex power delivered by a threephase source or extracted by a three-phase load.



Apply power-factor compensation.



Calculate power quantities based on wattmeter measurements.

Houses

Distribution substation 10:1 step down

Service wire 120 V / 240 V single phase Shops

Between the power generating station and a residence or shop, power is transferred across transmission lines in a form known as three-phase power. What is three-phase power and why is it used? The intent of the present chapter is to answer these questions and to provide the tools for analyzing three-phase networks.

“book” — 2015/5/4 — 7:23 — page 567 — #2

567

Overview How does the electrical power generated by a power station get distributed to users with diverse voltage-level and powerconsumption requirements? The requirements of a large factory are very different from those of a building or a single residence. The power distribution network—often called the power grid— uses above- or underground transmission lines to transfer power between different locations, and employs step-up and stepdown transformers to change voltage levels at specific nodes in the grid. As explained later in this chapter, at a typical large electrical station, power is generated by spinning a large electromagnet past three separate stationary coils arranged evenly around a circular tube. The energy used to spin the electromagnet may come from a hydroelectric dam, diesel or gas engine, or a steam turbine driven by burning coal, oil, or

All voltage values are rms

natural gas, or generated by a nuclear reactor. By spinning the electromagnet at 60 revolutions per second, magnetic induction generates an ac voltage across the terminals of each of the three coils. The three induced voltages have the same amplitude and frequency (60 Hz), but their phase angles are staggered by 360◦ /3 = 120◦ between any two of them. Hence, the power generated by this arrangement is called three-phase. In the power-distribution model shown in Fig. 10-1, the nuclear power station uses large transformers to step-up each of the coil voltages from its initial level to 480,000 V (rms). This is done before distributing the power across the grid. The conversion changes the voltage level, but not the amount of power made available by the transmission line. As noted in Section 7-10, when a transformer steps up the voltage by the turns ratio N2 /N1 , it simultaneously steps down the current by the same ratio. By stepping up the voltage at the power

Large factory

Large factory

Transmission lines 48,000 V 3-phase 480,000 V 3-phase

Primary substation

Nuclear power station

Transmission lines 48,000 V

Intermediate substation

Transmission lines 48,000 V 3-phase

20:1 step-down pole transformer

Service wire 120 V / 240 V single phase Houses

Distribution lines 16,000 V 3-phase

Distribution lines 4,800 V 3-phase

Distribution lines 4,800 V Distribution substation 10:1 step down

se

3-pha

Building

Underground distribution lines 4,800 V 3-phase

Service wire 120 V / 240 V single phase Shops

Building Figure 10-1: Typical electrical power grid.

“book” — 2015/5/4 — 7:23 — page 568 — #3

568

CHAPTER 10 THREE-PHASE CIRCUITS

Power lines

3600 turns 180 turns

Residential service cable Residence

120 V ac 4800 V 120 V ac Connected to Earth ground TV

Transformer

Computer

Stove

120 V ac Neutral

All voltage values are rms

240 V ac

120 V ac Earth ground

Vacuum cleaner

Light bulb

Figure 10-2: A 4800 V rms single-phase ac source connected to a residential user through a 20 : 1 step-down transformer.

station to a very high level, the current that flows along the transmission line gets reduced significantly. Since the power 2 R, where R is loss in the transmission line is Ploss = Irms the total resistance of the transmission line, the voltage upconversion step serves to reduce the power loss by several orders of magnitude. The power grid includes several substations designed to convert power from transmission to distribution. This is accomplished through the use of step-down transformers and a bus circuit that can split the power into multiple directions. For a single residence, a center-tapped pole transformer is used to step-down one of the three-phase lines to a level manageable by household applicances (Fig. 10-2). The power carried to the house from the transformer is called three-wire single phase, with the middle wire assuming the role of the neutral wire. It is single phase because the two 120 V rms voltages at the secondary side of the transformer have the same frequency and phase.

Instead of providing a return path to the generating station through an actual wire, the earth ground is used to provide the feedback path for electrons. This is accomplished by using a cable to connect the middle wire at the transformer output to ground (Fig. 10-2). Use of the ground cable is a safety measure to prevent charging up machinery to dangerous levels, as well as for the discharging of high-voltage events (like lightning) and to prevent current-related heating of wires due to unbalanced loads.

10-1

Balanced Three-Phase Generators

Figure 10-3(a) is a representative cross-sectional view of a typical three-phase ac generator. The generator consists of a rotating electromagnet, called the rotor, and three separate stationary coils distributed evenly around a circular tube called

“book” — 2015/5/4 — 7:23 — page 569 — #4

10-1

BALANCED THREE-PHASE GENERATORS

569

1

1

n ω S

3

2

V12

+ V _1 + V _2 + V _3

Rotor

N

_

+

Stator

3

_ +

V31

V23

+

_

2 (a) Three-phase generator

V3

V

Im √2 VYs

120o V1 −120o

υ1

υ2

υ3

t

Re

V2 (b) Phasor voltages V1 to V3 in the complex plane

(c) Voltage waveforms

Figure 10-3: Three-phase ac generator and associated voltage waveforms. the stator. The rotor is spun around by a turbine or some other external force. The three coils are arranged 120◦ apart over the circumference of the stator. As the electromagnet rotates, its magnetic field induces a sinusoidal voltage at the terminals of each of the three coils. If the coils are identical in shape and number of turns, the three induced phasor voltages, V1 to V3 , will all have the same amplitude and their timedomain counterparts, υ1 (t) to υ3 (t), will vary sinusoidally at the same frequency f = ω/2π , where ω is the angular rotation frequency of the rotor. However, because the coils are physically distributed 120◦ apart, the voltages induced in adjacent coils will be delayed in time and shifted in phase by 120◦ relative to one another. By designating the common terminal n as the neutral (ground) terminal with Vn = 0 and selecting V1 in Fig. 10-3(a) as the reference voltage with zero phase, the phase of V2 will be either 120◦ or −120◦ , relative to the phase of V1 , depending on the relative directions of the two windings. If

all windings are the same, which usually is the case, common practice is to adopt a positive (123) phase sequence in which case the phase of V2 follows behind that of V1 by 120◦ and the phase of V3 follows behind that of V2 by 120◦ . Hence, the phases of V1 , V2 , and V3 are 0, −120◦ , and −240◦ (or equivalently, +120◦ ) (Fig. 10-3(b)), and their waveforms are shifted in time accordingly (Fig. 10-3(c)). We will refer to this arrangement as a balanced three-phaseY-source configuration with a positive phase sequence: Y-Source Configuration V1 = VYs 0◦ , V2 = VYs −120◦ , V3 = VYs

−240◦

,

(10.1)

“book” — 2015/5/4 — 7:23 — page 570 — #5

570

CHAPTER 10 THREE-PHASE CIRCUITS

with rms magnitude VYs . Voltages V1 to V3 are called the phase voltages of the Y-source configuration, and they all have the same magnitude, VYs .

1 I1 + _

 Throughout the remaining material in this chapter, all magnitudes of phasor voltages and currents will denote rms values. 

3

√ √ υ1 (t) = Re[ 2 V1 ej ωt ] = 2 VYs cos ωt, √ where we included the factor 2 because VYs was specified as an rms value. Similarly, for the other two phasor voltage sources √ υ2 (t) = 2 VYs cos(ωt − 120◦ ) √ 2 VYs cos(ωt − 240◦ ).

A common alternative to the Y-source configuration is the -source configuration shown in Fig. 10-4(b). Note that the -source configuration does not have a neutral wire. The

+

_

−120o

2

(a) Y-source configuration 1 V31 = √3 VYs

150o

+

I31 3

I12

+

_

+

The time-domain counterpart of phasor voltage V1 of the Y-source configuration is

V2 = VYs

I2

I3

_

 In reality, associated with each source is a complex source impedance, but the three source impedances usually are ignored because their values are much smaller than those of the impedances of the loads connected to the generator circuit. 

υ3 (t) =

−240o

(10.2)

which can be verified numerically by inserting Eq. (10.1) into Eq. (10.2) or graphically by summing the three vectors in Fig. 10-3(b). In the wiring configuration of Fig. 10-3(a), which is redrawn diagrammatically in Fig. 10-4(a), the three voltage sources share neutral terminal n and a common wire called the neutral wire. This configuration, which may or may not include the neutral wire, is the most common in North America.

and

V3 = VYs

0o

neutral wire

_

V1 + V2 + V3 = 0,

n

+

In a positive phase sequence, clockwise rotation between sources in the complex plane (Fig. 10-3(b)) entails an incremental phase shift of −120◦ . In a negative phase sequence, the phase-shift increment is +120◦ . We should note that for a balanced three-phase source

V1 = VYs

_ V12 = √3 VYs

30o

I23 2

V23 = √3 VYs

−90o

(b) Δ-source configuration Figure 10-4: Y- and -source configurations, with VYs = rms value of the phase-voltage magnitude of the Y-source. The rms √ magnitude of the -source phase voltages is 3 VYs . relationships between voltages V12 , V23 , and V31 of the  configuration and the three voltages of the Y configuration are -Source Configuration V12 = V1 − V2 = VYs 0◦ − VYs −120◦ √ = 3 VYs 30◦ = Vs 30◦ , V23 = V2 − V3 = Vs

−90◦

(10.3)

,

150◦

V31 = V3 − V1 = Vs √ with Vs = 3 VYs .

√ We note that the magnitude of the -source, Vs , is 3 times larger than that of the Y-source, and the phases of the  sources are 30◦ ahead of theirY counterparts (see Fig. E10.1 of Exercise 10-1).

“book” — 2015/5/4 — 7:23 — page 571 — #6

10-1

BALANCED THREE-PHASE GENERATORS

571

 It is important to remember that the phase angle of a phasor is defined relative to a reference. In Fig. 10-4, we assigned V1 a phase angle of zero, and all of the other voltages are defined relative to V1 . Had its phase angle been φ instead, the phases of all of the other voltages would have had to be adjusted accordingly (V2 = VYs φ − 120◦ , etc.). 

Exercise 10-1: Superimpose onto Fig. 10-4(b) the three

source voltages of the  configuration. Answer:

Im V3

V31

V12 120o

30o

30o

Example 10-1: Y- Sources

If in a balanced -source configuration with a positive phase sequence, υ12 (t) = 440 cos(120π t + 45◦ ) V, determine υ1 (t), υ2 (t), and υ3 (t) of the equivalent Y-source configuration.

V2

30o

Solution: The rms phasor counterpart of υ12 (t) is

V23

440 V12 = √ 45◦ V (rms). 2

Figure E10.1

(See

From Fig. 10-4, we observe that V1 can be √ obtained from V12 by reducing the magnitude of the latter by 3 and decrementing its phase by 30◦ : V12 440 V1 = √ −30◦ = √ √ 45◦ − 30◦ = 179.63 15◦ 3 2 3 Hence, υ1 (t) =

√ 2 × 179.63 cos(120π t + 15◦ )

−120o

Re

V1

)

Exercise 10-2: Given a balanced -source configuration

with a positive phase sequence and V12 = 208 45◦ V (rms), determine (a) phase voltages V23 and V31 , and (b) (rms). V1 , V2 , and V3 of the equivalent Y-source configuration. Answer: (a) V23 = 208 −75◦ V (rms),

V31 = 208 −195◦ V (rms), (b) V1 = 120 15◦ V (rms), V2 = 120 −105◦ V (rms), V3 = 120 −225◦ V (rms). (See C3 ) )

= 254.03 cos(120π t + 15◦ ) V. By extension, υ2 (t) = 254.03 cos(120π t − 105◦ ) V, υ3 (t) = 254.03 cos(120π t + 135◦ ) V.

Exercise 10-3: Show graphically √ why the phase

magnitude of V12 of the -source is 3 times larger than the phase magnitude of the Y-source. Answer:

Im

Concept Question 10-1: Why are the voltage sources of a three-phase source separated by 120◦ increments in phase? (See )

Concept Question 10-2: In a Y-source configuration, each voltage source is measured across one of the generator’s three coils, with one terminal serving as a common neutral terminal for all three. How are the sources of a  configuration measured? (See )

−V2 V∆s 30˚ VYs

V1

V2 Figure E10.3 (See

)

V12 = V1 − V2

Re

“book” — 2015/5/4 — 7:23 — page 572 — #7

572

CHAPTER 10 THREE-PHASE CIRCUITS

10-2 Source-Load Configurations The three-phase generator is equivalent to three separate singlephase generators, each one of which can be connected to a separate load. Transmission of three-phase power from the generator to loads is more efficient than three separate, singlephase transmissions. A three-phase -source configuration uses 3 transmission wires and a Y-source configuration uses 3 or 4 wires (the neutral wire is not always used), whereas 6 wires are required to support three separate single-phase transmissions. The three loads connected to the three sources may be arranged in either a Y- or a -load configuration (with the  configuration being more common). Hence, the source-load connections may assume any one of four possible combinations: Y-Y, Y-, -Y, and - (Fig. 10-5).

Currents I1 , I2 , I3 I12 , I23 , I31 IL1 , IL2 , IL3

Iab , Ibc , Ica Impedances

ZTL1 , ZTL2 , ZTL3 Zn Za , Zb , Zc

 By applying equivalent-circuit transformations, any one of the four connection configurations can be converted into any of the other three. 

10-2.1 Y and  Notation In view of the several sources, currents, and impedances involved in the four source-load topologies, it will prove helpful to summarize our notation. We will be guided by the circuits in Fig. 10-5. Nodes 1, 2, 3 a, b, c n and N

nodes in source circuit nodes in load circuit neutral node in source and load (Y-Y configuration only)

phase currents in Y-source configuration phase currents in -source configuration line currents through transmission lines, same as phase currents Ia , Ib , Ic of Y-load phase currents of -load

Zab , Zbc , Zca

transmission-line impedances impedance of neutral line (Y-Y configuration only) impedances of Y-load configuration (for balanced load Za = Zb = Zc = ZY ) impedances of -load configuration (for balanced load Zab = Zbc = Zca = Z )

 We should note that the term line voltage is short-hand for line-to-line voltage at the point of use (i.e., at the load).  In the present context, the line voltages are the voltages between transmission-line pairs, namely the voltages between nodes a and b, b and c, and c and a in Fig. 10-5. The associated line currents are the currents flowing through the transmission lines. The terms phase voltages and phase currents usually refer to the voltages across and currents through the load impedances, but often they are also used in connection with the source circuit. To avoid confusion, it is best to specify whether it is the source circuit or the load circuit that the phase voltages and currents belong to.

Voltages V 1 , V 2 , V3 V12 , V23 , V31 Vn = 0 Vab , Vbc , Vca

VaN , VbN , VcN

phase voltages of Y-source, with rms magnitude VYs phase voltages of -source,√ with rms magnitude Vs = 3 VYs neutral node in Y-source line voltages (voltages between transmission-line pairs), with rms magnitude VL ; same as phase voltages of -load phase voltages of Y-load

10-2.2

Balanced Conditions

• In view of how the voltage sources are induced, it is safe to assume that the three Y or  sources are balanced, meaning that they have the same amplitude and frequency and their phases are separated by 120◦ increments. • The load circuit, however, may or may not be balanced. The load circuit is considered balanced if all of its impedances are the same. Three-phase motors are one such example.

“book” — 2015/5/4 — 7:23 — page 573 — #8

10-2

SOURCE-LOAD CONFIGURATIONS

IL1

1 I1

Ia

N

+

_

V2

I2

IL3 2

3

Ib

b

Load Phase Voltages VaN, VbN, VcN

a Iab

Ica Zca

Load Phase Currents Iab, Ibc, Ica

Zab

n V2

3

Y-source

+

_

Ic

c

+ _ V1

_

+

I3

Zb

IL1

1

V3

Zc

Y-load IL2 Transmission lines (a) Y-Y configuration

Y-source

I1

Load Phase Currents Ia, Ib, Ic (same as line currents IL1, IL2, and IL3)

Za

n

+

I3

a

+ _ V1

_

V3

573

I2

IL3 2

c

IL2

Ibc

Zbc

Load Phase Voltages Vab, Vbc, Vca

b

∆-load

Transmission lines (b) Y-∆ configuration

Figure 10-5: The three-phase source and load circuits can be connected in four possible arrangements: Y-Y, Y-, -Y, and -. In each arrangement, the source and load circuits are connected via transmission lines carrying line currents IL1 , IL2 , and IL3 . [Parts (c) and (d) follow on the next page.]

• A network is said to be balanced if its source voltages are balanced and if it has identical transmission lines and identical loads. The techniques presented in Chapters 7–9 are more than sufficient to analyze any three-phase circuit, for any combination of source and load configurations. Concept Question 10-3: What is a balanced source?

Balanced load? Balanced network? (See

)

Concept Question 10-4: Which line(s) do the line

voltages refer to? (See

)

Concept Question 10-5: For which load configuration

are (a) the phase voltages the same as the line voltages? and (b) the phase currents the same as the line currents? (See )

“book” — 2015/5/4 — 7:23 — page 574 — #9

574

CHAPTER 10 THREE-PHASE CIRCUITS

IL1

1 I12 V31 +_

Ia

ZTL

+ V _ 12

N IL3

I23

+

2

Zc

Zb

Ic

Ib

b

c

_

V23 IL2 Transmission lines

∆-source

a Ica

+ V _ 12

+

IL3

I23 2

c

Load Phase Currents Iab, Ibc, Ica

Zab

Ibc

_

V23 ∆-source

Iab

Zca

I31 3

(c) ∆-Y configuration

ZTL

I12

Load Phase Voltages VaN, VbN, VcN

Y-load

IL1

1

V31 +_

Load Phase Currents Ia, Ib, Ic (same as line currents IL1, IL2, and IL3)

Za

I31 3

a

Zbc

∆-load

IL2 Transmission lines

b

Load Phase Voltages Vab, Vbc, Vca (same as source voltages if ZTL is negligible)

(d) ∆-∆ configuration (Fig. 10-5 continued)

10-3 Y-Y Configuration In the Y-Y configuration depicted in Fig. 10-6, the Y-load network is connected to the Y-source circuit through four wires (transmission lines). The transmission lines, which may or may not be identical, are characterized by impedances ZTL1 , ZTL2 , and ZTL3 , connecting sources V1 through V3 to loads Za to Zc . In addition, a fourth transmission line of impedance Zn connects node n of the source configuration to node N of the load configuration.

(a) Develop a node voltage equation for VN , the voltage at node N , with node n treated as (the ground) reference. (b) Determine line currents iL1 (t) to iL3 (t) and in (t) for a balanced network, given that VYs = 120 V (rms), f = 60 Hz, ZTL1 = ZTL2 = ZTL3 = (1 + j 1) , and Za = Zb = Zc = (29 + j 9) . (c) Determine line voltages υab (t), υbc (t), and υca (t). Solution: (a) Relative to node n (i.e., with Vn = 0), the node equation at node N is:

Example 10-2: Balanced Y-Y Network

With reference to the network shown in Fig. 10-6:

In − IL1 − IL2 − IL3 = 0,

(10.4)

“book” — 2015/5/4 — 7:23 — page 575 — #10

10-3 Y-Y CONFIGURATION

575

V1

a

+ _

+ _

Za

_

_

Vab

In

Zn

n Vn = 0 V3

IL1

ZTL1

1

Vca VN N

Zb

Zc

+

+

V2

3

IL2

ZTL2

2

IL3

ZTL3

Y-source

_+

_

Vbc

b

Transmission

+ c

Y-load

Figure 10-6: Three-phase Y source connected to a Y load circuit via transmission lines.

and

or equivalently VN (V1 − VN ) (V2 − VN ) (V3 − VN ) − − − = 0. (10.5) Zn ZTL1 + Za ZTL2 + Zb ZTL3 + Zc (b) For a balanced network, the denominators of terms 2 to 4 in Eq. (10.5) become identical, which we shall denote as Z0 : Z0 = ZTL1 + Za = (1 + j 1) + (29 + j 9) = (30 + j 10) . (10.6) The node voltage equation given by Eq. (10.5) simplifies to:   1 V1 + V2 + V3 3 VN = + . (10.7) Zn Z0 Z0 According to Eq. (10.2), V1 + V2 + V3 = 0. Hence, VN = 0

for

a

balanced

(balanced network).

source,

(10.8)

V3 − VN 120e−j 240 ◦ = = 3.80ej 101.6 A (rms). Z0 30 + j 10

The numerical values of phasor currents IL1 to IL3 are in rms, so to obtain their corresponding time-domain expressions, we √ need to multiply them by 2: √ iL1 (t) = Re[ 2 IL1 ej ωt ] = 5.37 cos(2πf t − 18.4◦ ) A, √ iL2 (t) = Re[ 2IL2 ej ωt ] = 5.37 cos(2πf t − 138.4◦ ) A, and

√ iL3 (t) = Re[ 2IL3 ej ωt ] = 5.37 cos(2πf t + 101.6◦ ) A,

with f = 60 Hz. (c) Vab = IL1 Za − IL2 Zb

Consequently,



VN In = = 0, (10.9) Zn V1 − VN 120 ◦ IL1 = = 3.80e−j 18.4 A (rms), = Z0 30 + j 10 ◦

IL2



IL3 =

V2 − VN 120e−j 120 ◦ = = = 3.80e−j 138.4 A (rms), Z0 30 + j 10



= (3.80e−j 18.4 − 3.80e−j 138.4 )(29 + j 9) ◦

= 200ej 28.8 V (rms). Similarly, ◦

Vbc = IL2 Zb − IL3 Zc = 200e−j 91.2 V (rms) ◦

Vca = IL3 Zc − IL1 Za = 200ej 148.8 V (rms).

“book” — 2015/5/4 — 7:23 — page 576 — #11

576

CHAPTER 10 THREE-PHASE CIRCUITS

The corresponding time-domain line voltages are

a



υab (t) = 282.2 cos(2πf t + 28.8 ) V, υbc (t) = 282.2 cos(2πf t − 91.2◦ ) V,

ZY

and υca (t) = 282.2 cos(2πf t + 148.8◦ ) V.

ZY

Concept Question 10-6: What is the magnitude of In in

a balanced Y-Y configuration (Fig. 10-6)? (See

ZY

c

b

)

Z∆ = 3ZY Exercise 10-4: Were we to repeat Example 10-2, but with the transmission-line impedances set to zero, which of the following line-current quantities will change and which will remain the same: (a) amplitudes, (b) absolute phases, and (c) phases relative to each other?

a

Answer: (a) Amplitudes will change, (b) absolute

Z∆

phases will change, but (c) relative phases will continue to be 120◦ apart (between pairs). (See )

10-4 Balanced Networks A three-phase network is balanced if it has a balanced threephase source, transmission lines with identical impedances, and a balanced three-phase load (with equal impedances). The inherent symmetry associated with a balanced network allows us to simplify the circuit analysis considerably.

10-4.1 Transformation Between Balanced Sources

c

Z∆

Z∆

b

Figure 10-7: Y- transformation for balanced load circuits.

 Transformation of (V1 , V2 , V3 ) of the Y-source into -source involves multiplication (V12 , V23 , V31 ) of the √ of the magnitudes by 3 and advancing the phases by 30◦ . 

A balanced, three-phase Y-source with a positive phase sequence is completely specified by two parameters: (a) VYs , the rms magnitude of the phase voltages and (b) φ1 , the phase of V1 . That is, ⎧ ⎪ ⎨V1 = VYs φ1 , (10.10) Balanced Y-source = V2 = VYs φ1 − 120◦ , ⎪ ⎩ V3 = VYs φ1 + 120◦ . According to our earlier discussion in connection with Fig. 10-4, the Y-source can be transformed into an equivalent -source, ⎧ √ ◦ ⎪ ⎨V12 = V1 × √3 30 , (10.11) Balanced -source = V23 = V2 × 3 30◦ , √ ⎪ ⎩ V31 = V3 × 3 30◦ .

Z = 3ZY .

(10.12)

“book” — 2015/5/4 — 7:23 — page 577 — #12

TECHNOLOGY BRIEF 25: MINIATURIZED ENERGY HARVESTING

577

Technology Brief 25 Miniaturized Energy Harvesting Energy is present in our environment in many forms. One can think of energy as a quantity that measures the ability of a system to do some amount of work on its environment (or other systems). For example, moving objects possess kinetic energy, objects within gravity wells possess potential energy, and charged particles within an electric field possess electrical energy. Real systems often transduce energy from one form to another as part of normal operation: photodiodes (solar cells) convert electromagnetic waves (light) into the movement of charge particles in a potential field (thus providing a system like your phone with the ability to do work), sensors often transduce chemical energy into electrical energy to measure the state of a system, an electrostatic actuator might transduce electrical energy into mechanical energy to perform mechanical work, etc. There is useful energy present all around you, in the gentle mechanical vibrations moving through your building, in the radio frequency waves generated by radio emitters, and even in the heat that your body or your car engine emits. As computation and communication technologies miniaturize and become ever-more pervasive, many everyday computing objects require less and less power to operate. A typical circa-2012 laptop might consume 50–100 W (joules per second) during normal operation, a smartphone might consume 0.5–4 W (depending on what the user is doing, whether the radio is on, etc.), and a good low power wristwatch might consume 10 μW down to ∼100 nW. As power consumption decreases for some functions, it turns out that there is, in many cases, just enough energy in the environment to power these systems. This is often known as energy harvesting or energy scavenging. Below, we’ll look at some interesting devices that have been built to scavenge energy from the environment to power everyday systems. A great many scavenging systems have been built in recent years, so we’ll focus on general classes of scavenging. It is also important to note that the line between power scavenging and conventional power generation can become blurred: is a normal solar cell scavenging light to produce power? Sure! Is a wind turbine scavenging wind power to produce electricity? Of course. The idea is to focus on technologies that convert very small sources of power which, in the past, were often too small to be useful or were ignored.

∆V

∆T ∆V = (s1 − s2) ∆T Figure TF25-1: The most common class of thermoelectric materials operate according to the Seebeck effect. When two conductors are joined at one end and exposed to a temperature gradient (T ), a potential difference (V ) is measurable across the free ends of the two conductors. The relationship between V and T depends on (s1 −s2 ), where s1 and s2 are the Seebeck coefficients of the two materials. The Seebeck coefficient is a material-specific property thet depends on the molecular structure of the material. This potential difference can be used to drive a current through an external load and thus do work. Interestingly, the effect can be run in reverse—known as the Peltier effect—such that an applied voltage can be used to create a temperature difference. This is the basis of cryogenic cooling systems.

Thermoelectric Almost every system that does useful work also produces heat. This “waste heat” is often exhausted to the environment but, since time immemorial, humans have also used heat to do work. Steam engines, internal combustion engines, the turbine systems at power plants and thousands of other heat engines extract useful energy from a heat source. Modern, top of the line power plants contain combined heat and power (CHP) systems that internally recover waste heat to increase efficiency. A number of miniaturized technologies have been explored for scavenging tiny amounts of waste heat. Although these efforts have included making tiny heat engines, fundamental physical limitations have so far prevented successful scaling down of mechanical heat engines down to the millimeter. A different approach is to use materials that convert heat directly to electrical power; thermoelectric materials are in this category. In the early 19th century Thomas Johann Seebeck observed that a voltage was induced when two dissimilar conductors were placed in a thermal gradient (Fig. TF25-1). Thermoelectric materials are used extensively to sense temperature. More recently,

“book” — 2015/5/4 — 7:23 — page 578 — #13

578

TECHNOLOGY BRIEF 25: MINIATURIZED ENERGY HARVESTING

Figure TF25-2: (top) A tiny micromechanical harvesting system developed at the University of Michigan. It occupies 27 mm3 and can harvest > 200 μW delivered at 1.85 V when exposed to 1.5 g’s of acceleration when vibrating at or near ∼155 Hz.

researchers have built tiny thermoelectric systems that scavenge power from small thermal gradients (such as is found between the surface of your skin and the environment); the small temperature difference (usually 1◦ –5 ◦ C) and the low efficiency of existing thermoelectric materials has limited this to low scavenged energy densities ( 0. • Load is capacitive if QT < 0.

Next, if we set Zab = Zbc = Zca = Z and φab = φbc = φca = φ in Eqs. (10.35) and (10.38), and then subtract P1 from P2 , we end up with

• Load is resistive if QT = 0. Concept Question 10-12: A wattmeter uses a current coil and a voltage coil. How should the two coils be connected relative to the source and load of a single-phase circuit? (See )

Concept Question 10-13: The two-wattmeter method

P2 − P1 =

VL2 Z

[cos(φ − 60◦ ) + cos φ − cos φ − cos(φ + 60◦ )]

=

VL2 [cos(φ − 60◦ ) − cos(φ + 60◦ )]. Z

Applying the trigonometric identity cos(x ± y) = cos x cos y ∓ sin x sin y,

can provide a measurement of what power quantity in a three-phase network? Is the method constrained to balanced networks? (See ) Exercise 10-8: When used on a balanced three-phase

load, the two-wattmeter method provided measurements P1 = 4,800 W and P2 = 10,200 W. What is the total complex power ST of the load? Answer: ST = (15000 + j 5400) VA. (See

)

“book” — 2015/5/4 — 7:23 — page 595 — #30

10-7

POWER MEASUREMENT IN THREE-PHASE CIRCUITS

595

Summary Concepts • In a balanced three-phase source with a positive phase sequence, the three sources have identical voltage magnitudes and frequency, but their phase angles are shifted clockwise by −120◦ increments. • Three-phase networks can assume four configurations: Y-Y, Y-, -Y, and -, and each can be transformed into any of the other three. • A balanced three-phase network can be subdivided into three independent, single-phase circuits.

• The total instantaneous power supplied by a balanced three-phase source is a constant (not a function of time), thereby rendering smooth power delivery to a balanced three-phase load, such as a three-phase motor. • The two-wattmeter method provides a measurement of the total real power PT consumed by any three-phase load, whether balanced or not. Moreover, if the load is balanced, the method also provides a measurement of the reactive power QT .

Mathematical and Physical Models Balanced -Source V12 = Vs φ1 + 30◦ V23 = Vs φ1 − 90◦ V31 = Vs φ1 + 150◦ √ Vs = 3 VYs V12 + V23 + V31 = 0

Balanced Y-Source V1 = VYs φ1 V2 = VYs φ1 − 120◦ V3 = VYs φ1 + 120◦ V1 + V 2 + V 3 = 0 Y

 Balanced Loads Z = 3ZY

Two-Wattmeter Measurement PT = P (any load) √1 + P2 QT = 3 (P2 − P1 ) (Balanced Y- or -load)

Important Terms -load configuration -source configuration average real power balanced balanced load balanced network balanced source balanced three-phase Y-source configuration bus circuit center-tapped pole transformer complex power current coil

Total Complex √ Power ST = 3 VL IL φY √ ST = 3 VL IL φ

(balanced Y-load) (balanced -load)

Total Instantaneous Power PT (t) = 3VYL IYL cos φY PT (t) = 3VL IL cos φ

(balanced Y-network) (balanced -network)

Provide definitions or explain the meaning of the following terms: digital wattmeter line current line voltage line-to-line voltage magnitude magnitude of the -source negative phase sequence network neutral node neutral terminal neutral wire phase current phase magnitude phase voltage

positive (123) phase sequence power factor power grid reactive power real power rms magnitude of the phase current rms magnitude of the phase voltages rotor single-phase generators stator step-down transformer step-up transformer

three-phase three-phase ac generator three-wire single phase total average power total reactive power transformer transmission line two-wattmeter method voltage coil wattmeter Y-load configuration Y-source configuration Y-Y, Y-, -Y, - configurations

“book” — 2015/5/4 — 7:23 — page 596 — #31

596

CHAPTER 10 THREE-PHASE CIRCUITS

PROBLEMS

j44 Ω

ZTL1

a

18 Ω

ZTL2

b

18 Ω −j16 Ω

ZTL3

c

14 Ω

Section 10-1: Balanced Three-Phase Generators

Three-phase source

10.1 For each of the following groups of sources, determine if the three sources constitute a balanced source, and if it is, determine if it has a positive or negative phase sequence. (a) υa (t) = 169.7 cos(377t + 15◦ ) V υb (t) = 169.7 cos(377t − 105◦ ) V υc (t) = 169.7 sin(377t − 135◦ ) V (b) υa (t) = 311 cos(ωt − 12◦ ) V υb (t) = 311 cos(ωt + 108◦ ) V υc (t) = 311 cos(ωt + 228◦ ) V

10.8 In the network of Fig. P10.8, Za = Zb = Zc = (25 + j 5) . Determine the line currents.

10.2 In a balanced Y-source with a positive phase sequence, V1 = (103.92 − j 60) V (rms). Determine: (a) V2 and V3 and (b) V12 , V23 , and V31 of the equivalent -source configuration, all in polar form.

10.5 In a balanced -source with a positive phase sequence, V23 = (56.94+j 212.5) V (rms). Determine υ12 (t), υ23 (t), and υ31 (t). Assume f = 60 Hz. 10.6 In a balanced Y-source with a positive phase sequence, V1 = (16.93 + j 46.51) V (rms). Determine υ12 (t), υ23 (t), and υ31 (t) of the equivalent -source configuration. Assume f = 60 Hz.

Sections 10-2 to 10-4: Configurations and Networks *10.7 In the circuit of Fig. P10.7, VaN = 65 47◦ V (rms), VbN = 60 15◦ V (rms), VcN = 45 −86◦ V (rms), and f = 60 Hz. Calculate υab (t), υbc (t), and υca (t).

200 n 200

Vrms

+

−95o

Vrms

+

145o

_

*10.4 In a balanced Y-source, V2 = 100 V (rms) and V3 = −100 V (rms). Is the phase sequence positive or negative? Also, determine V1 .

25o

_

60◦

200

_

10.3 In a balanced -source with a positive phase sequence, υ12 (t) = 240 cos(120πt −20◦ ) V. Determine (a) υ23 (t), υ31 (t), and (b) υ1 (t), υ2 (t), and υ3 (t) of the equivalent Y-source configuration.

Answer(s) available in Appendix G.

j33 Ω

Figure P10.7: Circuit for Problem 10.7.

(c) V1 = 140 −140◦ V V2 = 114 −20◦ V V3 = 124 100◦ V



N

+

Vrms

1

IL1

a

Za

2

IL2

b

Zb

3

IL3

c

Zc

N

Figure P10.8: (Problems 10.8 to 10.12.)

10.9 In the network of Fig. P10.8, Za = (10 + j 2) , Zb = 6 , and Zc = (8 + j 1) . Determine the line currents. 10.10 Repeat Problem 10.8 after inserting transmission-line impedances ZTL = (3 + j 1)  between nodes 1 and a, 2 and b, and 3 and c. *10.11 Repeat Problem 10.9 after inserting transmission-line impedances ZTL = (3 + j 1)  between nodes 1 and a, 2 and b, and 3 and c. 10.12 Repeat Problem 10.9 after adding a zero-impedance wire between nodes n and N . Also, determine the current through it. 10.13 Apply single-phase equivalency to determine the line currents in the Y- network shown in Fig. P10.13. The load impedances are Zab = Zbc = Zca = (25 + j 5) .

“book” — 2015/5/4 — 7:23 — page 597 — #32

PROBLEMS

120

597

0

_

120 n

+

−120

_

120

Vrms

Vrms

+

120

+

Vrms

IL1

1



2



IL2

3



IL3

in order to use single-phase equivalency to determine the three line currents. Transmission-line impedances are ignored.

a

10.16 Given a balanced - network with V12 = 440 0◦ V (rms) connected in a positive phase sequence, and Z = (6 − j 2) , apply the necessary transformation in order to use single-phase equivalency to determine the three line currents. Transmission-line impedances are ignored.

Zab b

Zca

Zbc

*10.17 Determine IL1 in the network of Fig. P10.17. (Hint: This is a balanced network. If you apply the correct transformations, the solution simplifies considerably.)

c

_

Figure P10.13: (Problems 10.13 to 10.14.)

10.18 The network shown in Fig. P10.18 consists of a balanced 120 V (rms)Y-source connected in a positive phase sequence, lossless transmission lines, and an unbalanced -load with Zab = 14 , Zbc = (6 − j 2) , and Zca = (24 + j 6) . Assign V1 a phase angle of 0◦ .

*10.14 Determine the line currents in the network of Fig. P10.13, given that Zab = 10 , Zbc = 5 , and Zca = (10 − j 5) .

(a) Determine the phase voltages at the load: Vab , Vbc , and Vca .

10.15 Given a balanced -Y network with V12 = 440 0◦ V (rms) connected in a positive phase sequence, and ZY = (10 − j 2) , apply the necessary transformation

+ _ 60 _ +

−30

2 + _ 60

−150 Vrms

(4 − j2) Ω (6 + j1) Ω

(6 + j1) Ω



3

a (6 + j1) Ω (4 − j2) Ω b

Vrms 2Ω

(4 − j2) Ω

c

Figure P10.17: (Problem 10.17.)

IL1

1 I1

V3

Ica

Iab

Zca

Zab

n V2

_

3

+

I3

a

+ _ V1

+

90 Vrms

_

60

IL1



1

(b) Determine the phase currents: Iab , Ibc , and Ica .

I2 2

IL3

c

IL2 Figure P10.18: (Problem 10.18.)

Ibc

Zbc

b

“book” — 2015/5/4 — 7:23 — page 598 — #33

598

CHAPTER 10 THREE-PHASE CIRCUITS

IL1

1

a

I12 V31 +_

Ica

+ V _ 12

Zca

I31 _

3

+

Iab

IL3

I23 2

Zab

Ibc

c

Zbc

b

V23 IL2 Figure P10.19: (Problem 10.19.)

*10.20 Determine In in the circuit of Fig. P10.20. All sources are rms. Sections 10-5 to 10-7: Power 10.21 For the network in Fig. P10.21, (a) generate three single-phase equivalent circuits and (b) determine the complex power supplied by the three-phase source. 10.22 For the network in Fig. P10.22, (a) generate three single-phase equivalent circuits and (b) determine the complex power supplied by the three-phase source. *10.23 Determine the complex power supplied by the source in the network of Fig. P10.23. 10.24 A balanced Y-load is supplied by a three-phase generator at a line voltage of 416 V (rms). If the real power

0.2 + j1 + o _ 100 0 100 n _ _

10.19 The network shown in Fig. P10.19 consists of a balanced 120 V (rms) -source connected in a positive phase sequence, lossless transmission lines, and an unbalanced -load with Zab = (6 + j 4) , Zbc = 4 , and Zca = 12 . Assign V12 a phase angle of 0◦ . (a) Determine the phase voltages at the load: Vab , Vbc , and Vca . (b) Determine the phase currents: Iab , Ibc , and Ica . (c) Determine the line currents: IL1 , IL2 , and IL3 . (d) Determine the total power absorbed by the load, given that the power absorbed by an impedance Z due to the flow of current I through it is P = |I|2 Re[Z], where I is rms.

IL1

+ 100

−120o

0.2 + j1

−240o

IL2 j80 Ω 70 Ω

I2 0.2 + j1

I3

j14 Ω 84 Ω

I1

+

(c) Determine the line currents: IL1 , IL2 , and IL3 . (d) Determine the total power absorbed by the load, given that the power absorbed by an impedance Z due to the flow of current I through it is P = |I|2 Re[Z], where I is rms.

0.2 + j1

−j18 Ω

N

IL3 45 Ω

In

Figure P10.20: Circuit for Problem 10.20.

absorbed by the load is 6 kW at a power factor 0.7 lagging, determine ZY and the magnitude of the line current. 10.25 A balanced -load is supplied by a three-phase generator at a line voltage of 208 V (rms). If the complex power extracted by the load is (8 + j 4) kVA, determine Z and the magnitude of the line current. *10.26 For the network in Fig. P10.23, determine (a) the average real power supplied by the three-phase source and (b) what fraction of it is absorbed by the three-phase load. 10.27 Determine the complex power extracted by the load in Fig. P10.27. Also determine the power factor of the overall load circuit as seen by the source.

“book” — 2015/5/4 — 7:23 — page 599 — #34

PROBLEMS

599

0

220

_

220



+

−120

n

_

220

Vrms

5Ω + _ 200

Vrms



+

120

(8 + j2) Ω

Vrms



+

(8 + j2) Ω

N

(8 + j2) Ω

_ 0

Vrms

+

_

−120

220 n

+

_

120

220

Vrms

Vrms

+

(6 − j2) Ω (6 − j2) Ω (6 − j2) Ω 2Ω

_

0

_

220 n

+

−120

_

220

Vrms

Vrms

+

120

+

Vrms

2Ω (8 + j2) Ω 2Ω

+ _ 200

−120

10 Ω

j4 Ω

Vrms 8Ω

j2 Ω

10.31 A 240 V (rms), 60 Hz Y-source is connected to a balanced three-phase Y-load by four wires, one of which is the neutral wire. If the load is 400 kVA at pf old = 0.6 lagging, what size capacitors should be added to change the power factor to pf new = 0.95 lagging? *10.32 A 416 V (rms), 60 Hz Y-source is connected to a balanced three-phase Y-load by four wires, one of which is the neutral wire. If the load is 800 kW at pf old = 0.75 lagging, what size capacitors should be added to change the power factor to pf new = 0.95 lagging?

Figure P10.22: (Problem 10.22.)

220

_ Vrms +

*10.30 A 416 V balanced three-phase source supports four loads connected in parallel. Each load is itself a balanced threephase load. Determine the line current and the power factor at the source, given that load 1 is 12 kVA at pf 1 = 0.7 leading, load 2 is 18 kVA at pf 2 = 0.9 lagging, load 3 is 6 kW at pf 3 = 1, and load 4 is 24 kVA at pf 4 = 0.7 lagging.





120

Figure P10.27: (Problem 10.27.)

Figure P10.21: (Problem 10.21.)

220

200

0 Vrms

−j2 Ω

(10 + j1) Ω (6 − j4) Ω



_

Figure P10.23: (Problems 10.23 and 10.26.)

*10.28 In the three-phase network of Fig. P10.28, ZTL = 4 , Z1 = (10 − j 4) , and ZY2 = (5 + j 2) . Determine the line currents. 10.29 A 208 V (rms) balanced three-phase source supports two loads connected in parallel. Each load is itself a balanced three-phase load. Determine the line current, given that load 1 is 12 kVA at pf 1 = 0.7 leading and load 2 is 18 kVA at pf 2 = 0.9 lagging.

10.33 A balanced three-phase source with a line voltage of 208 V (rms) is connected to a three-phase motor designed as a balanced Y-load. The powers measured using the twowattmeter method are P1 = 800 W and P2 = 300 W. Determine the impedances of the motor and its power factor. *10.34 Determine the power readings of the two wattmeters shown in the circuit of Fig. P10.34 given that Z = (10 + j 6)  and the amplitudes of the voltage sources are rms. 10.35 Determine the power readings of the two wattmeters shown in the circuit of Fig. P10.35 given that ZY = (15 − j 5) . *10.36 Determine the power readings of the two wattmeters shown in the circuit of Fig. P10.36 given that Z = (10 + j 6)  and the amplitudes of the voltage sources are rms. 10.37 Repeat Problem 10.36 after replacing the balanced -load with a balanced Y-load with ZY = (10 + j 6) .

“book” — 2015/5/4 — 7:23 — page 600 — #35

600

CHAPTER 10 THREE-PHASE CIRCUITS

ZTL

120 V

+

+

_

280

_

280

_ 280

+

IL1

a

ZY2

Z∆1

0 V

ZTL

IL2

ZTL

IL2

ZY2 b

−120 V

Z∆1

ZY 2 c

Z∆1

Figure P10.28: Circuit for Problem 10.28.

+ _

+ _

P1

P1

+ _ 180

30

+ _

ZY

+ _ Z∆

180

30 V

+ _ ZY

Z∆ _

120

Z∆

+

180

120 V

_

180

+

+ _ + _

+ _

P2

Figure P10.35: Circuit for Problem 10.35.

+ _

Potpourri Questions

+ _

How are the Seebeck and Peltier effects related?

480 −30o V _

10.40 Why do power stations use large-size capacitors and inductors?

P1

Z∆ Z∆

480 −150o V _

10.41 What is a stator coil?

+ _

+

10.39

ZY

P2

Figure P10.34: Circuit for Problem 10.34.

10.38 Name three types of energy harvesting and describe the energy conversion process in each case.

+ _

+

480

P2

90o V

Z∆

+ _ + _

Figure P10.36: Circuit for Problem 10.36.

“book” — 2015/5/4 — 7:25 — page 601 — #1

11

CHAPTER

Magnetically Coupled Circuits

Contents 11-1 TB27 11-2 11-3 11-4 11-5

Overview, 602 Magnetic Coupling, 602 Magnetic Resonance Imaging (MRI), 608 Transformers, 611 Energy Considerations, 615 Ideal Transformers, 617 Three-Phase Transformers, 619 Summary, 622 Problems, 623

Objectives Learn to: 

Incorporate mutual coupling in magnetically coupled circuits.



Analyze circuits containing magnetically coupled coils.



Relate input to output voltages, currents, and impedances for magnetically coupled transformers, including ideal transformers and three-phase transformers.

Magnetic flux i1

i2

+ υ1

_

N1

+ N2

υ2

_ Primary port Secondary port When two physically unconnected inductors are in close proximity to one another, current flow through one of them induces a magnetically coupled voltage across the other one. Magnetic coupling may be intentional or not. Highly coupled voltage transformers used in power distribution networks are an example of intentional coupling. If the coupling between two coils in a circuit is unintentional but significant, its effects should be incorporated into the analysis of the circuit.

“book” — 2015/5/4 — 7:25 — page 602 — #2

602

CHAPTER 11

MAGNETICALLY COUPLED CIRCUITS

Overview Voltage transformers are used in many electrical systems, including power supply circuits (Section 7-9) and power distribution networks (Chapter 10). Whereas resistors, capacitors, and inductors are one-port, two-terminal devices, a transformer is a two-port (four-terminal) device with a primary port and a secondary port. Coupling of energy between the two ports is realized through a shared magnetic field, without the need for direct contact between them. Transformers are part of a family of devices and circuits called magnetically coupled circuits, whose operation relies on magnetic coupling rather than current conduction. We begin this chapter by examining the voltage and current relationships between the primary and secondary ports of a coupled two-coil system. We did so previously in Section 7-9.1, but then our treatment was limited to the special case of the ideal transformer with perfect coupling. In this more comprehensive examination, we introduce the concepts of mutual inductance, equivalent circuits, and impedance transformations, and we learn how three-phase transformers are configured to step-up or step-down voltage levels in three-phase power circuits.

i1(t)

+

~ + −

Φ12

Φ11

+

υ1

υ2

_

_ N1 turns N2 turns

(a) Current i1 induces Φ11 and Φ12, which induces υ2 i2(t)

+

Φ21

Φ22

+

υ1

υ2

_

_

~+−

N1 turns N2 turns (b) Current i2 induces Φ22 and Φ21, which induces υ1 Figure 11-1: Magnetically coupled coils.

11-1 Magnetic Coupling Magnetic coupling can occur between any two inductors in close proximity of one another. Current flow through the coils of one of the inductors induces a mutual inductance voltage across the other inductor, and vice versa. The induction process is described in terms of a mutual inductance, measured in henrys (H), that depends on the degree of magnetic coupling between the two inductors, which in turn depends on their physical shapes, orientations relative to one another, spacing between them, and the magnetic permeability μ of the medium between them. Mutual inductance may be intentional or not. It is key to the operation of highly coupled transformers used for stepping up and stepping down voltage levels. On the other hand, mutual inductance between two inductors, transmission lines, or wires in a certain circuit may be totally unintentional, as well as unavoidable. In that case, we should learn how to account for the voltages induced by the mutual inductance and how to incorporate them in the analysis of the circuit. The two magnetically coupled coils in Fig. 11-1(a) have N1 turns on primary port 1 and N2 turns on primary port 2. Port 1 is connected to a source that causes current i1 (t) to flow through coil 1, which generates magnetic flux 11 linking coil 1 alone and flux 12 linking both coils.

 Magnetic fluxes form closed loops because the magnetic field lines that emerge from one end of the primary coil will also flow back in at the other end of the coil. The direction of the magnetic field is dictated by the direction of the current in the coil: if the four fingers of the right hand point in the direction of the current, the thumb will point in the direction of the magnetic field. Alternatively, if the four fingers wrap in the direction of the magnetic flux, the thumb will point in the direction of the current. This is known as the right hand rule.  The total magnetic flux through coil 1 is 1 = 11 + 12 .

(11.1)

Magnetic flux linkage 1 is defined as the total flux linking all N1 turns of coil 1, 1 = N1 1 .

(11.2a)

For coil 2, the flux linking coil 1 to coil 2 is 12 , and the corresponding 2 is 2 = N2 12 .

(11.2b)

Self inductance L1 of coil 1 is defined as the ratio of the magnetic flux linkage 1 to the current i1 responsible for

“book” — 2015/5/4 — 7:25 — page 603 — #3

11-1

MAGNETIC COUPLING

603

inducing 1 ,

1 L1 = , (11.3) i1 and the voltage induced across inductor L1 is di1 d1 = L1 . (11.4a) υ1 = dt dt By analogy, 2 in coil 2 induces voltage υ2 , with d12 d2 = N2 . (11.4b) υ2 = dt dt Both υ1 and υ2 are induced by di1 /dt. In the case of υ1 , the link is self inductance L1 , as given by Eq. (11.4a). To establish an analogous relationship between di1 /dt and υ2 , we rewrite Eq. (11.4b) as d12 di1 υ2 = N2 × . (11.4c) di1 dt Next, we define the mutual inductance M21 , as d12 , (11.5) M21 = N2 di1 and the expression for υ2 becomes υ2 = ±M21

di1 . dt

(11.6)

 Subscripts 21 refer to the fact that M21 is the inductance of coil 2 due to the magnetic field induced by current i1 .  The expression in Eq. (11.6) includes a (±) on the righthand side. This is because mutual inductance M21 is a positive quantity measured in henrys (H), whereas υ2 may be positive or negative, depending on the direction of the winding in coil 2 relative to the direction of the winding in coil 1. For the specific winding directions shown in Fig. 11-1(a), the appropriate sign is (+). If we were to reverse the roles of coils 1 and 2, by connecting the source to coil 2 instead of to coil 1, thereby causing current i2 to flow through coil 2, as depicted in Fig. 11-1(b), we would end up with the following expressions for υ1 and υ2 : di2 (11.7a) υ1 = ±M12 dt and di2 υ2 = L2 . (11.7b) dt  Because the coupled coils constitute a linear system, energy considerations (Section 11-3) require that M12 = M21 = M, where M is now called the mutual inductance between the two coils. 

The ambiguity between the (+) and (−) signs in Eqs. (11.6) and (11.7a) is resolved through the use of a standard dot convention based on the directions of the two windings. For a specific direction of i1 (left-hand side Fig. 11-2), the polarity of υ2 depends on whether the dots are on the same or opposite terminals of the windings and whether i1 enters coil 1 at its dotted or undotted terminal.  In a two-coil magnetically coupled system, if current enters the first one at its dotted terminal, the polarity of the mutual-inductance voltage induced across the second coil is positive at its dotted terminal. The polarity of the induced voltage is reversed if the current in the first coil enters at the undotted terminal. Moreover, reciprocity applies: current in the second coil induces a mutualinductance voltage across the first one in accordance with the same dot convention.  This dot convention covers all combinations of current directions and dot locations outlined in Fig. 11-2. Finally, if we generalize to the configuration shown in Fig. 11-3(a) in which currents flow through both coils simultaneously, voltage υ1 will contain two components, one due to self-inductance of coil 1 and another due to the mutual inductance between the two coils. That is, υ1 will be the sum of Eqs. (11.4a) and (11.7a), and similarly, υ2 becomes the sum of Eqs. (11.6) and (11.7b). Specifically: For dots on same ends and currents entering coils at dotted ends (Fig. 11-3(a)):

υ1 = L1

di2 di1 +M dt dt

(11.8a)

di1 di2 +M . dt dt

(11.8b)

and υ2 = L2

For dots on opposite ends but current entering coils at same ends (Fig. 11-3(b)):

υ1 = L1

di1 di2 −M dt dt

(11.9a)

di1 di2 −M . dt dt

(11.9b)

and υ2 = L2

“book” — 2015/5/4 — 7:25 — page 604 — #4

604

CHAPTER 11

i1

MAGNETICALLY COUPLED CIRCUITS

M L1

L2

+ di1 υ2 = M dt _

+

di2 υ1 = M dt _

L2

L1

(a) i1

(e)

M L1

i2

M L2

+ di1 υ2 = −M dt _

+

di2 υ1 = −M dt L1 _

(b)

L2 (f)

M L1

i2

M

M L2

+ di1 υ2 = −M dt _

+

di2 υ1 = −M dt L1 _

L2 i2

i1 (c)

(g)

M L1

M L2

+ di1 υ2 = M dt _

+

di2 υ1 = M dt _

L1

L2 i2

i1 (d)

(h)

Figure 11-2: Dot convention for the mutual-inductance voltage induced in coil 2 by current i1 in coil 1, and vice versa.

Primary Side: υ1

Sign Convention

(1) The self-induced component L1 di1 /dt of υ1 is assigned a (+) sign if: (a) the (+) polarity of υ1 is defined at the dotted terminal, and i1 enters coil 1 at the dotted terminal, or (b) the (+) polarity of υ1 is defined at the undotted terminal, and i1 enters coil 1 at the undotted terminal. Otherwise, L1 di1 /dt is assigned a (−) sign. (2) The mutually induced component M di2 /dt of υ1 is assigned a (+) sign if: (a) the (+) polarity of υ1 is defined at the dotted terminal, and i2 enters coil 2 at the dotted terminal, or (b) the (+) polarity of υ1 is defined at the undotted terminal, and i2 enters coil 2 at the undotted terminal. Otherwise, M di2 /dt is assigned a (−) sign. Secondary Side: υ2 The same sign convention applies from the perspective of the secondary side: simply replace “primary” with “secondary,” subscript 1 with 2, and vice versa.

“book” — 2015/5/4 — 7:25 — page 605 — #5

11-1

MAGNETIC COUPLING

605

i1

R1 υ s1

i2

M

+

+ _

di1 di2 υ1 = L1 dt + M dt _ Change + to − if i1 is CCW

R2

+

L2

L1

Change + to − if i2 is CW

+ _

di2 di1 υ2 = L2 dt + M dt _

Change + to − if i2 is CW

υs2

Change + to − if ii is CCW

(a) Dots on same ends

i1

R1 υ s1

+ _

i2

M

+

di1 di2 υ1 = L1 dt − M dt _ Change + to − if i1 is CCW

R2

+

L1

Change − to + if i2 is CW

L2

+ _

di2 di1 υ2 = L2 dt − M dt _

Change + to − if i2 is CW

υs2

Change − to + if ii is CCW

(b) Dots on opposite ends Figure 11-3: Polarities of voltage components for clockwise (CW) and counterclockwise (CCW) current directions.

Example 11-1: 1 kHz Circuit

R1 Determine load current iL (t) in the circuit of Fig. 11-4(a), given that υs (t) = 10 cos(2π × 103 t) (V), R1 = 5 , C1 = C2 = 10 μF, L1 = 1 mH, L2 = 3 mH, M = 0.5 mH, and RL = 20 . Solution: We start by transforming the ac circuit from the time domain to the phasor domain (Fig. 11-4(b)). Thus, υs (t) becomes phasor voltage Vs , C gets transformed into an impedance ZC = −j/ωC, and L gets transformed into an impedance ZL = j ωL. The angular frequency is

υs(t)

+ _

C1

C2

M

iL L1

L2

RL

(a) Time domain R1 −j/ωC1 Vs

+ _

I1

jωL1

jωM

−j/ωC2 IL jωL2

I2

RL

ω = 2πf = 2π × 10 rad/s. 3

Denoting I1 and I2 as the mesh currents in the two loops, both defined with clockwise directions, the mesh-current equations are   j + j ωL1 I1 − j ωMI2 = 0 (11.10a) −Vs + R1 − ωC and

 −j ωMI1 + j ωL2 −



j + RL I2 = 0, ωC

(11.10b)

(b) Phasor domain Figure 11-4: Circuit of Example 11-1.

where C = C1 = C2 . Note that the polarity of the last term in Eq. (11.10a) is negative because, in accordance with the convention shown in Fig. 11-2(g), the winding dots are on the same end in Fig. 11-4(b) but I2 enters the undotted terminal of

“book” — 2015/5/4 — 7:25 — page 606 — #6

606

CHAPTER 11

coil 2. Simultaneous solution of the two equations for IL = I2 gives j ωMVs  j R1 + j ωL1 − ωC RL + j ωL2 −

IL = 

j ωC



Substitution of the specified values leads to IL = 139.5e

j 142.2◦

The first mesh equation is −Vs + R1 I1 + V1 − V2 = 0,

.

+ ω2 M 2 (11.11)

MAGNETICALLY COUPLED CIRCUITS

or equivalently, (R1 + j ωL1 + j ωL2 − j 2ωM)I1 − (j ωL2 − j ωM)I2 = Vs . (11.13a) Similarly, for the second mesh,

mA,

and its time-domain equivalent is V2 + RL I2 = 0,

iL (t) = Re[IL ej ωt ] = 139.5 cos(2π × 103 t + 142.2◦ ) mA. or equivalently, Example 11-2: Coupled Inductors

−(j ωL2 − j ωM)I1 + (RL + j ωL2 )I2 = 0.

The circuit in Fig. 11-5 has the following element values: ◦ Vs = 30ej 60 V, L1 = 10 mH, L2 = 30 mH, R1 = 5 , RL = 10 , and the ac source operates at 60 Hz. The circuit layout is such that inductors L1 and L2 experience a relatively small mutual inductance M. Determine the average power delivered to the load RL for (a) M = 4 mH, (b) M = 1 mH, and (c) M = 0. Solution: Before we apply mesh analysis, let us determine V1 and V2 across the two inductors. Voltage V1 consists of two terms, j ωL1 I1 due to current I1 entering at the (+) terminal of V1 , and j ωM(I2 − I1 ) due to current (I2 − I1 ) through L2 . The polarity of the second term is governed by the dot convention: if current enters a coil at its dotted terminal, the polarity of the mutual-inductance voltage induced across the second coil is positive at its dotted terminal. In the present case, (I2 −I1 ) enters L2 at its dotted terminal, so the voltage it induces across L1 is positive at the dotted terminal of L1 . Hence, V1 = j ωL1 I1 + j ωM(I2 − I1 ).

(11.13b)

(a) M = 4 mH Upon replacing R1 , RL , L1 , L2 with their specified values, setting M = 4 mH, and multiplying inductances by ω = 2πf = 2π × 60 = 377 rad/s, matrix solution of the two equations gives ◦

I2 = 1.657ej 63.1 A,

(11.14)

and according to Eq. (8.3), the corresponding average power absorbed by RL is PL =

1 1 |I2 |2 RL = (1.657)2 × 10 = 13.73 W. 2 2

(11.15)

(b) M = 1 mH Repetition of the process with M = 1 mH gives

(11.12a)



I2 = 1.64ej 62.15 A

(11.16)

PL = 13.45 W.

(11.17)

Application of the same rule to L2 gives V2 = j ωL2 (I2 − I1 ) + j ωMI1 .

R1

Vs

+ _

(11.12b)

jωL1

(c) M = 0

+ V1 __ V2

I1 jωM

and

+

In the absence of mutual coupling between the two coils,

jωL2

I2

Figure 11-5: Circuit of Example 11-2.

RL



I2 = 1.635ej 62.03 A

(11.18)

PL = 13.372 W.

(11.19)

and

“book” — 2015/5/4 — 7:25 — page 607 — #7

11-1

MAGNETIC COUPLING

607

Error

when connected to the same voltage source Vs . For the twoinductor circuit in Fig. 11-6(b),

Ignoring M altogether would incur an error in PL of % error =

PL (@M = 4 mH) − PL (@M = 0) × 100 PL (@M = 4 mH)

= 2.61%,

when true M is 4 mH,

% error = 0.61%,

when true M is 1 mH.

and

I1 = I2 = I, and while I1 enters L1 at its dotted terminal, I2 enters L2 at its undotted terminal. While guided by Fig. 11-3(b), application of the dot convention to the loop in Fig. 11-6(b) gives V1 = j ωL1 I1 − j ωMI2 = j ω(L1 − M)I and V2 = j ωL2 I2 − j ωMI1 . = j ω(L2 − M)I.

Example 11-3: Equivalent Inductance

For the circuit in Fig. 11-6(a), obtain an expression for the equivalent inductance, Leq , defined such that it would exhibit the same i-υ characteristic at nodes (a, b) as the actual circuit. Solution: Equivalency means that circuits in Figs. 11-6(b) and (c) will have the same current I flowing through both loops

At terminals (a, b), Vs = V1 + V2 = j ω(L1 + L2 − 2M)I. For the circuit in circuit Fig. 11-6(c), Vs = j ωLeq I.

a

L1

Equivalency leads to Leq = L1 + L2 − 2M.

(a)

Leq

L2

M b

Concept Question 11-1: What determines the polarity

of the mutual inductance voltage? Summarize the rules of the dot convention. (See )

I (b)

Vs

a I1

+ V1 _ +

+ _ b I

(c)

Vs

L1

M

V2

_

I2

Concept Question 11-2: What factors determine how strong or weak the magnetic coupling is between two coils? (See )

L2 Exercise 11-1: Repeat Example 11-1 after moving the dot location on the side of L2 from the top end of the coil to the bottom.

a

Answer: iL(t) = 139.5 cos(2π × 103t − 37.8◦) mA.

(See

+ _

)

Leq b

Figure 11-6: Finding Leq of two series-coupled inductors (Example 11-3).

Exercise 11-2: Repeat Example 11-3 for the two in-series

inductors in Fig. 11-6(a), but with the dot location on L2 being on the top end. Answer: Leq = L1 + L2 + 2M. (See

)

“book” — 2015/5/4 — 7:25 — page 608 — #8

608

TECHNOLOGY BRIEF 27: MAGNETIC RESONANCE IMAGING (MRI)

Technology Brief 27 Magnetic Resonance Imaging (MRI)

Precession N

Magnetic resonance imaging (MRI), also called nuclear magnetic resonance (NMR), is a powerful medical imaging tool that provides extremely detailed 2-D and 3-D images of the body, an example of which is shown in Fig. TF27-1. MRI is particularly useful for imaging soft tissues (organs, ligaments, spinal column, arteries and veins, etc.). Unlike X-ray, which uses very high frequency ionizing radiation, MRI uses lower frequency magnetic and radio frequency fields, which are non-ionizing and do not damage cells. Since MRI’s early demonstration in the 1970s, its applications have burgeoned, and research is continually opening up new and improved MRI techniques. MRI utilizes the fact that the bulk of the human body contains water and every water molecule has a permanent magnetic dipole moment, which means that a water molecule behaves like a small magnet. The hydrogen atoms in the (H2 O) water molecule have a natural spin associated with them, and because they are weakly charged, this spin (charges moving in a circle) creates a magnetic field as shown by the S-to-N arrow in Fig. TF27-2. Thus, the water molecule acts like a weak bar magnet with North and South poles. Normally, these spins are randomly aligned in the body, but if a strong external magnetic field is applied, they all line up with the applied magnetic field. Almost exactly half line up in the N-S direction, and almost exactly half in the opposite SN direction. This dc magnetic field will just hold them in place for the rest of the MRI scan time. But you can never have exactly half of the spins in each of the two directions, so a few extra spins (about 9 out of 2 million for a typical 1.5 tesla magnet) always

Figure TF27-1: MRI scan of the head.

Spin

Applied external magnetic field S

Figure TF27-2: The spin of the charged hydrogen atoms in H2 O produces a magnetic field. show up in neither the N-S or S-N configuration. These extra outlier molecules are the ones the MRI scanner actually uses to create the image. A second magnetic field is then applied, but unlike the strong dc magnetic field imposed to hold the rest of the molecules in place, this magnetic field is a radiofrequency (RF) pulsed signal, at a particular frequency for which the outlier molecules are known to resonate strongly (this is the resonance part of MRI). This natural resonant frequency is called the Larmor frequency, and it depends on the chemical makeup, density, and structure of the tissue, thus allowing MRI to distinguish different tissues, identify and detect chemical composition, and even determine the status (such as inflammation) of various tissues. When the RF pulse turns on, the outlier molecules align with that magnetic field. When the pulse turns off, they relax back to their original state. As they relax, their spin precesses (becomes tipped like a toy top slowing down) and produces yet another magnetic field, which returns to and is picked up by the same coil that produced the original RF signal. Now let’s look at the MRI machine and the hardware (Fig. TF27-3) that makes this all happen. The large applied magnetic field is produced by a large superconducting electromagnet. A typical medical MRI scanner is 1.5 teslas (1 tesla = 10k gauss). By comparison, a strong refrigerator magnet is 100 gauss, and the Earth’s magnetic field at its surface is around 0.5 gauss. The superconducting magnet is cooled with liquid nitrogen down to the point where its resistance is virtually zero (see Technology Brief 3 on superconductivity), so that the current and hence magnetic field can be maximized. These magnets weigh several tons and cost hundreds of thousands of dollars a year in electricity and liquid nitrogen to keep them running. They take weeks to cool down enough to reach superconductivity, and days to ramp up the current to produce their large magnetic field.

“book” — 2015/5/4 — 7:25 — page 609 — #9

TECHNOLOGY BRIEF 27: MAGNETIC RESONANCE IMAGING (MRI)

609

Superconducting Magnet

Figure TF27-3: MRI scanner geometry. The patient is slid into the bore (hole) in the center of the magnet, where the field is strong, and also quite uniform. Because the body is “nonmagnetic” (μr = 1), this strong magnetic field does not move or hurt the person, although it does polarize (line up) the spins in his/her hydrogen atoms. If the person has any metal inside of him/her (such as implantable medical devices, artificial joints, or bone repair surgeries), this can preclude the use of MRI for that patient. It is also important to keep all metal (oxygen tanks, wheelchairs, pens, clipboards, etc.) away from the magnet, or it can be pulled irretrievably and dangerously into the bore. The magnetic field is so strong that these materials cannot be removed without slowly (days or weeks) reducing the current in the electromagnet to turn it off. The MRI scans the body in slices, like a loaf of bread. The slice being scanned is adjusted by creating a very small gradient in the RF field using yet another set of coils, the gradient coils shown in Fig. TF27-4. The current in these coils is ramped up and turned on and off very quickly to move through hundreds of scan slices quickly.When the current in a coil decreases, the magnetic field decreases too. The energy stored in this magnetic field has to go somewhere—it is returned to the circuit, creating a voltage spike at the source driving the coil.The voltage spike must be controlled by managing the current decay. In addition, this decreasing magnetic field creates another current that tries to oppose the change (you may have learned about Lenz’s law in physics), making it impossible to instantly turn off a large magnetic field. Another interesting (and often very noisy!) effect is seen in these gradient coils. The coils themselves vibrate from their strong magnetic field, which creates a constant hum and often loud

Figure TF27-4: MRI scanner gradient coils are used to adjust which slice the scanner is imaging.

thumps and even crashing noises, so patients receiving MRI scans generally wear earplugs to block the noise. Now let’s look at the RF coils that transmit and receive the RF pulses. RF coils come in many different designs, shapes, and sizes as shown in Fig. TF27-5. The coil must be large enough to surround the region being imaged (the head, torso, knee, etc.). The most common coils for whole-body or whole-head imaging are the birdcage coil shown in part (a) of the figure and the planar or surface coils shown in part (b). The signal of interest is determined by how much of the RF energy gets from the coil to the feature of interest, and back to the coil. Coils such as the birdcage coil are designed to have a field as constant as possible over the head, for instance, with no particular focus on the optic nerve. So, the signal (S) from the head is relatively large but the signal from the optic nerve is relatively small. The electrical noise (N) picked up by the receiver is generated by the entire field of view of the coil. So for imaging the head, the signal-to-noise ratio (SNR), given by S/N, is relatively large, but for the optic nerve, the signal is lower while the noise is the same, so SNR is smaller and not good enough to provide an accurate measurement of the signal. Much of the research on MRI coils today is therefore focused on development of coils such as the one in part (c) of Fig. TF27-5, which focus the RF energy on a specific feature of interest (in this case, the optic nerve) thus increasing the signal (S) in the SNR and often decreasing the noise (N) as well. The effect of SNR on image quality is seen in Fig. TF27-6.

“book” — 2015/5/4 — 7:25 — page 610 — #10

610

TECHNOLOGY BRIEF 27: MAGNETIC RESONANCE IMAGING (MRI)

(a) Birdcage coil

(b) Surface coil

(c)

Figure TF27-5: Coils used in MRI: (a) birdcage coil; (b) surface coil (one on the front and another on the back of the body). (Credit: Emilee Minalga.)

(a) Low SNR

(b) High SNR

Figure TF27-6: Imaging of the optic nerve (seen just below the eyeballs) with (a) low SNR and (b) high SNR. (Photo courtesy Robb Merrill.)

“book” — 2015/5/4 — 7:25 — page 611 — #11

11-2 TRANSFORMERS

611

11-2 Transformers 11-2.1

Transformer core i1

Coupling Coefficient

To couple magnetic flux between two coils, the coils may be wound around a common core (Fig. 11-7(a)), on two separate arms of a rectangular core (Fig. 11-7(b)), or in any other arrangement conducive to having a significant fraction of the magnetic flux generated by each coil shared with the other. The coupling coefficient k defines the degree of magnetic coupling between the coils, with 0 ≤ k ≤ 1. For a loosely coupled pair of coils, k < 0.5; for tightly coupled coils, k > 0.5; and for perfectly coupled coils, k = 1. The magnitude of k depends on the physical geometry of the two-coil configuration and the magnetic permeability μ of the core material. A transformer is said to be linear if μ of its core material is a constant, independent of the magnitude of the currents flowing through the coils (and hence, the strength of the induced magnetic field). 

+ _

N1

υ1

i2

+ N2

k=

12 12 = . 11 + 12 1

(11.20a)

where 1 = 11 +12 . The perfectly coupled case corresponds to when the flux coupled to coil 2, namely 12 , is equal to the self-coupled flux 1 . Similarly, from the standpoint of coil 2,

(a) Cylindrical core

Magnetic flux i1

i2

+

+

N1

_

N2

21 21 = . 22 + 21 2

(11.20b)

υ2

_ Primary port Secondary port (b) Rectangular core Figure 11-7: Magnetically coupled coils.

Through energy considerations, k can be related to L1 , L2 , and M as k=√

M . L 1 L2

(11.21)

The mutual inductance M is a maximum when k = 1 (perfectly coupled transformer), M(max) =

k=

_

Magnetic flux

υ1 Most core materials, including air, wood, and ceramics, are nonferromagnetic, and their μ is approximately equal to μ0 , the permeability of free space. When nonferromagnetic materials are used for the common core around which the coils are wound, the magnitude of k depends entirely on how tightly coupled the two windings are. Such transformers are indeed linear, but the magnitude of k is seldom greater than 0.4. Increasing k requires the use of ferromagnetic cores, but the transformer becomes heavier in weight and its behavior becomes nonlinear. The degree of nonlinearity depends on the choice of materials. With certain types of purified iron, transformers can be designed to exhibit coupling coefficients approaching unity. As was noted earlier in connection with Fig. 11-1(a), current i1 , through coil 1 generates magnetic fluxes 11 through coil 1 and 12 through both coils 1 and 2. The coupling coefficient is given by

υ2

 L1 L2 .

(11.22)

(perfectly coupled transformer with k = 1)

“book” — 2015/5/4 — 7:25 — page 612 — #12

612

CHAPTER 11

+ _

Vs

R2 L2

L1

I1

We note that

M

R1

a

MAGNETICALLY COUPLED CIRCUITS

I2

ZR =

ZL

ω2 M 2 . impedance of secondary loop

b (a) Original circuit

In the absence of coupling between the two windings of the transformer (i.e., M = 0), Zin reduces to

a

Zin (M = 0) = R1 + j ωL1 .

+ _

Vs

I1

Zin

b (b) Equivalent circuit Figure 11-8: (a) Transformer circuit with coil resistors R1 and R2 , and (b) in terms of an equivalent input impedance Zin .

11-2.2

This is exactly what we expect for a series RL circuit connected to a source Vs . When M is not zero, the impedance of the secondary circuit, (R2 + j ωL2 + ZL ), becomes part of the input impedance of the primary circuit, enabled by the magnetic coupling represented by M. This dependence is akin to reflecting the impedance of the secondary circuit onto the primary circuit. The input and reflected impedances are related by

Input Impedance

In addition to the two coupled coils, a realistic transformer circuit should include two series resistors, R1 and R2 , to account for ohmic losses in the coils. The circuit shown in Fig. 11-8 reflects this reality by including resistor R1 on the side of coil 1 and resistor R2 on the side of coil 2. The circuit is driven by a voltage source Vs on the primary side and terminated in a complex load ZL on the secondary side. In terms of the designated mesh currents I1 and I2 , the KVL mesh equations are −Vs + (R1 + j ωL1 )I1 − j ωMI2 = 0

(11.23a)

−j ωMI1 + (R2 + j ωL2 + ZL )I2 = 0.

(11.23b)

Zin = Zin (M = 0) + ZR .

(11.26)

The expressions given by Eqs. (11.24) and (11.25) were derived for a transformer circuit in which the windings have dots on the same ends. Repetition of the process for windings whose dots are on opposite ends leads to the same results.  Zin depends on the degree of magnetic coupling, but not on whether the coupling is additive or subtractive. 

and From the standpoint of source Vs , the circuit to the right of terminals (a, b) can be represented by an equivalent input impedance Zin , as depicted in Fig. 11-8(b). By manipulating Eqs. (11.23) to eliminate I2 , we can generate the following expression for Zin : Zin =

Vs ω2 M 2 = (R1 + j ωL1 ) + I1 R2 + j ωL2 + ZL = (R1 + j ωL1 ) + ZR ,

(11.24)

where we define the second term as the reflected impedance ZR , namely ZR =

ω2 M 2 . R2 + j ωL2 + ZL

(11.25)

Example 11-4: Input Impedance

Determine current I1 in the circuit of Fig. 11-9. Solution: From the given circuit, we deduce that ωM = 2 . By analogy with Eq. (11.24), Zin is given by Zin = (3 − j 2 + j 5) +

22 ◦ = 4.2ej 42.2 . 6 + 4 − j 4 + j 20

Hence, ◦

I1 =

Vs 120ej 30 −j 12.2◦ = A. ◦ = 28.6e Zin 4.2ej 42.2

“book” — 2015/5/4 — 7:25 — page 613 — #13

11-2 TRANSFORMERS

613

I1 120

30o

V

+ _



a

−j2 Ω

Zin

−j4 Ω

j2 Ω

j5 Ω



j20 Ω



b Figure 11-9: Circuit of Example 11-4.

11-2.3

Equivalent Circuits

T-Equivalent Circuit

A circuit is said to be electrically equivalent to another if both exhibit the same I-V relationships at a specified set of terminals. For the transformer in Fig. 11-10(a), phase voltages V1 and V2 are related to I1 and I2 by V1 = j ωL1 I1 + j ωMI2

(11.27a)

V2 = j ωL2 I2 + j ωMI1 ,

(11.27b)

In anticipation of next steps, we have joined in Fig. 11-10(a) the negative terminals of V1 and V2 together, which has no impact on the operation of the transformer. Part (b) of the figure displays a proposed T-equivalent circuit whose element values Lx , Ly and Lz automatically incorporate the magnetic coupling present in the transformer coils, thereby avoiding the need to account for the mutual-inductance terms when writing KVL equations. The I-V matrix equation for the T-circuit (also called aY-circuit) is

and

which can be cast in matrix form as





    V1 j ωL1 j ωM I1 = . j ωM j ωL2 I2 V2

    I1 j ωLz V1 j ω(Lx + Lz ) = . (11.28) V2 j ωLz j ω(Ly + Lz ) I2

(11.27c)

(T-equivalent circuit)

(transformer) The transformer and its T-equivalent circuit exhibit the same I-V relationships if the four terms in the matrix of Eq. (11.27)

I1 V1

_

Lx

Ly

_

+ L1

L2

V2

_

Lz

(b) T-equivalent circuit

Lc

I1

I2

+ V1

I2

+

(a) Transformer

I1

M

+

+

V2

V1

_

I2

+ Lb

_

(c)

∏-equivalent

La

circuit

Figure 11-10: The transformer can be modeled in terms of T- or -equivalent circuits.

V2

_

“book” — 2015/5/4 — 7:25 — page 614 — #14

614

CHAPTER 11

are identical with their corresponding terms in the matrix of Eq. (11.28). Equalization of the two matrices leads to

MAGNETICALLY COUPLED CIRCUITS

If the transformer dots are located on opposite ends, M in Eq. (11.31) should be replaced with −M.

Transformer dots on same ends Lx = L1 − M,

(11.29a)

Ly = L2 − M,

(11.29b)

and Lz = M.

(11.29c)

Example 11-5: Equivalent Circuit

Use the T-equivalent circuit model to determine I1 in the circuit of Fig. 11-11. Solution: Use of Eq. (11.29) gives

Had the transformer dots been located on opposite ends, the two terms involving M in Eq. (11.27) would have been preceded by minus signs. Consequently, the element values of inductors Lx , Ly , and Lz would be

j ωLx = j ωL1 − j ωM = j 5 − j 2 = j 3 , j ωLy = j ωL2 − j ωM = j 20 − j 2 = j 18 , and j ωLz = j ωM = j 2 .

Transformer dots on opposite ends Lx = L1 + M,

(11.30a)

Ly = L2 + M,

(11.30b)

and Lz = −M.

(11.30c)

Even though a negative value for inductance Lz is not physically realizable, the mathematical equivalency holds nonetheless and the equivalent circuit is perfectly applicable.

The T-equivalent circuit is shown in Fig. 11-11(b). Application of the mesh analysis by-inspection method leads to the matrix equation 

(3 − j 2 + j 3 + j 2) −j 2 −j 2 (j 2 + j 18 − j 4 + 6 + 4)  ◦ 120ej 30 . = 0

  I1 I2

-Equivalent Circuit In some situations, it may be easier to analyze the larger circuit within which the transformer resides by replacing the transformer with a -equivalent circuit instead of the T-equivalent circuit. In such cases, we can use the model shown in Fig. 11-10(c). The expressions for La , Lb , and Lc can be obtained either by repeating the procedure we used for the T-equivalent circuit or by applying the Y- transformation equations given in Section 7-4.2. Either route leads to:

Its solution is ◦

I1 = 28.6e−j 12.2 A, which is identical with the answer obtained in Example 11-4 using the input impedance method. Concept Question 11-3: What does the coupling

coefficient represent? What is its range? (See

Transformer with dots on same ends

)

L1 L2 − M 2 , L1 − M

(11.31a)

L1 L2 − M 2 Lb = , L2 − M

(11.31b)

tance M related to L1 and L2 for a perfectly coupled transformer? (See )

(11.31c)

Concept Question 11-5: Why does the reflected impedance ZR bear that name? (See )

La =

Concept Question 11-4: How is the mutual induc-

and Lc =

L1 L2 − M 2 . M

“book” — 2015/5/4 — 7:25 — page 615 — #15

11-3

ENERGY CONSIDERATIONS

I1 120

30o

V

615

a



−j2 Ω

+ _

−j4 Ω

j2 Ω

j5 Ω



j20 Ω



b (a) Original circuit

I1 120

30o

V

a



+ _

−j2 Ω

j18 Ω −j4 Ω

j3 Ω jωLx jωLz

I1 b

jωLy j2 Ω



I2



(b) Equivalent circuit Figure 11-11: (a) Original circuit and (b) after replacing transformer with T-equivalent circuit.

Exercise 11-3: The expression for Zin given by

Eq. (11.25) was derived for the circuit in Fig. 11-8, in which both dots are on the upper end of the coils. What would the expression look like were the two dots located on opposite ends? Answer: The expression remains the same. (See

)

Exercise 11-4: What are the element values of the

-equivalent circuit of the transformer in Fig. 11-11(a)? Answer: j ωLa = j 32 , j ωLb = j 5.33 ,

j ωLc = j 48 . (See

)

For the circuit in Fig. 11-12(b), 1 1 1 Lx i12 + Ly i22 + Lz (i1 + i2 )2 2 2 2 1 1 1 = (L1 − M)i22 + (L2 − M)i22 + M(i1 + i2 )2 2 2 2 1 1 = L1 i12 + L2 i22 + Mi1 i2 . (11.33) 2 2

w(t) =

Equation (11.33) applies to transformers in which i1 and i2 both enter or both leave the dotted terminals, and both dotted terminals are on the same end (as in Fig. 11-12). Reversing the direction of either current or reversing the locations of the dots requires replacing M with −M.

11-3 Energy Considerations

Example 11-6: Magnetic Energy

Given that the transformer in Fig. 11-12(a)—with inductance L1 on the primary side, L2 on the secondary side, and mutual inductance M coupling the two coils—is equivalent to the T-equivalent circuit in Fig. 11-12(b), we can use the latter to compute the total amount of energy stored in the transformer for any specified values of currents i1 and i2 . According to Eq. (5.59), the magnetic energy stored in an inductor L due to the flow of current i through it is

In the circuit in Fig. 11-13, determine the magnetic energy stored in the transformer at t = 0, given that υs (t) = 12 cos(377t + 60◦ ) V. Solution: We start by replacing the transformer with its T-equivalent circuit and then transforming the new circuit to the phasor domain (Fig. 11-13(b)). Per Eq. (11.30), the values of Lx , Ly , and Lz are Lx = L1 + M = (10 + 6) mH = 16 mH,

1 w(t) = L i 2 (t) 2

(J).

(11.32)

Ly = L2 + M = (30 + 6) mH = 36 mH,

“book” — 2015/5/4 — 7:25 — page 616 — #16

616

CHAPTER 11 leads to

M

i1

i2

L1

j ωLx = j 377 × 16 × 10−3 = j 6 ,

L2

j ωLy = j 377 × 36 × 10−3 = j 13.57 , and

(a) Transformer

i1

MAGNETICALLY COUPLED CIRCUITS

j ωLz = j 377 × (−6 × 10−3 ) = −j 2.26 .

Lx = L1 − M

Ly = L2 − M

i2

Mesh analysis by inspection gives

i1 + i2



    j 60◦  (5 + j 6 − j 2.26) +j 2.26 I1 12e . = +j 2.26 (10 + j 13.57 − j 2.26) I2 0

M

(b) T-equivalent

Solution of the matrix equation for I1 and I2 leads to

Figure 11-12: Transformer and its T-equivalent circuit. Reversing the direction of either current or if dots are on opposite ends, M should be replaced with −M.



I1 = 1.91ej 26.06 A and ◦

I2 = 0.29e−j 112.5 A.

6 mH

5Ω υs

+ _

The time-domain equivalents are

10 mH

30 mH

10 Ω i1 (t) = 1.91 cos(377t + 26.06◦ ) A

(a) Original circuit

and i2 (t) = 0.29 cos(377t − 112.5◦ ) A.

j6 Ω



12

60o

V

+ _

I1

j13.57 Ω

10 Ω

−j2.26 Ω

The magnetic energy stored in the three inductors of the circuit in Fig. 11-13(b) at t = 0 is

I2



(b) Equivalent circuit in phasor domain

w(0) =

Figure 11-13: Circuits of Example 11-6.

=

and Lz = −M = −6 mH. Transforming the inductors to the phasor domain entails multiplying the inductance values by j ω = j 377 rad/s, which

 1 1 1 2 2 2 Lx i1 + Ly i2 + Lz (i1 − i2 ) 2 2 2 t=0

1 × 16 × 10−3 × (1.91 cos 26.06◦ )2 2 1 + × 36 × 10−3 × [0.29 cos(−112.5◦ )]2 2 1 + × (−6 × 10−3 ) 2 · [1.91 cos 26.06◦ − 0.29 cos(−112.5◦ )]2

= 13.7 mJ.

“book” — 2015/5/4 — 7:25 — page 617 — #17

11-4

IDEAL TRANSFORMERS

a + Vs _

a Zin b

I2

1:n

+ _

Zin b

+ Vs _

I1

617

c

+ _

V1

ZL

V2

N1 N2 (a) Dots on same ends

d

I1

c

I2

1:n

+ _

V1

For the ideal transformer with dots on the same ends (Fig. 11-14(a)), using Eqs. (11.34) and (11.35) in Eq. (11.27a) [while noting that the direction of I2 in Fig. 11-14(a) is opposite of that used to derive Eq. (11.27)] leads to

_

V1 = j ωL1 I1 − j ωMI2  = j ωL1 I1 − j ω L1 L2 I2

ZL

V +2

N1 N2 (b) Dots on opposite ends

 Note that the symbol for the ideal transformer includes two parallel lines between the coils, an indicator that the coils are wound around a ferromagnetic core with perfect coupling (k = 1). Also note that because μ of a ferromagnetic material is very large, so are L1 and L2 . 

d

= j ωL1 I1 − j ωL1 nI2 = j ωL1 (I1 − nI2 ).

Figure 11-14: Schematic symbol for an ideal transformer. Note the reversal of the voltage polarity and current direction when the dot location at the secondary is moved from the top end of the coil to the bottom end. For both configurations: V2 /V1 = N2 /N1 = n, I2 /I1 = N1 /N2 = 1/n.

Similarly, for Eq. (11.27b), V2 = j ωL2 I2 − j ωMI1 = [j ωL1 (I1 − nI2 )]n. Hence, V2 =n V1

(ideal transformer with dots on same ends).

(11.36)

11-4 Ideal Transformers An ideal transformer is characterized by lossless coils (R1 = R2 = 0) and a maximum coupling coefficient k = 1. The mutual inductance M is a maximum and given by Eq. (11.22) as  M(max) = L1 L2 .

(11.34)

 The transformer is called a step-up transformer (V2 > V1 ) when n > 1 and a step-down transformer (V2 < V1 ) when n < 1.  For a lossless transformer, complex power S1 supplied by its primary side must match complex power S2 absorbed by its secondary:

(ideal transformer) According to Eq. (5.52), the inductance L of a solenoid-shaped inductor is proportional to N 2 , where N is the number of turns wound around the core. Hence, for an ideal transformer with N1 turns on the primary side and N2 on the secondary, as depicted in Fig. 11-14, N2 L2 = 22 = n2 , L1 N1 where n = N2 /N1 is called the turns ratio.

(11.35)

S1 = S2 ,

(11.37)

V1 I1∗ = V2 I2∗ .

(11.38)

or equivalently In view of Eq. (11.36), it follows that I2 1 = I1 n

(ideal transformer with dots on same ends).

(11.39)

The expressions for the voltage and current ratios given by Eqs. (11.36) and (11.39) apply to the current directions, voltage polarities and dot locations indicated in both configurations of Fig. 11-14.

“book” — 2015/5/4 — 7:25 — page 618 — #18

618

11-4.1

CHAPTER 11

Input Impedance

Z1

At the input terminals in the transformer circuits of Fig. 11-14, source Vs views an input impedance Zin = V1 /I1 , and at the load end, ZL = V2 /I2 . The combination leads to Zin =

V1 1 V2 ZL = 2 = 2 . I1 n I2 n

 For the ideal transformer, the voltage, current, and impedance ratios are defined in terms of the turns ratio, independently of the values of L1 and L2 . 

11-4.2

a

I2

1:n

+ _

c

Z2

+ _

V1

+ _ Vs2

V2

b n = N2 /N1 d (a) Original circuit Z1

I1

2 a ZTh(ab) = Z2 /n

+ _

+ Vs1 _

+ VTh(ab) _ = Vs2 /n

V1

b (b) Thévenin equivalent of circuit to the right of terminals (a, b) ZTh(cd) = n2Z1

Equivalent Circuits

If the circuit connected to the transformer on its secondary side consists of only a passive impedance, the input-impedance equivalency is all we need to simplify the overall circuit, but what if the secondary side contains a voltage or current source? Also, by extension, what if we wish to view the circuit from the perspective of the secondary side, rather than that of the primary side? These questions are addressed by Example 11-7.

I1

+ Vs1 _

(11.40)

Thus, from the standpoint of the source Vs , the entire circuit to the right of terminals (a, b) is equivalent to input impedance Zin . We note that Zin is equal to (1/n2 ) of ZL .

MAGNETICALLY COUPLED CIRCUITS

VTh(cd) + = nVs1 _

Z2

c

+ _

+ _ Vs2

V2

d (c) Thévenin equivalent of circuit to the left of terminals (c, d) Figure 11-15: Th´evenin equivalent circuit of Example 11-7.

Example 11-7: Thevenin ´ Equivalent

For the circuit of Fig. 11-15, (a) obtain the Th´evenin equivalent of the circuit segment to the right of terminals (a, b), then (b) repeat the process for the circuit segment to the left of terminals (c, d). Solution: (a) The Th´evenin equivalent circuit shown in Fig. 11-15(b) consists of a Th´evenin voltage source VTh(ab) and a Th´evenin impedance ZTh(ab) . Our task is to relate them to the element values of the circuit to the right of terminals (a, b). The Th´evenin voltage is defined as the open-circuit voltage across terminals (a, b), which amounts to determining V1 after disconnecting the input circuit (Vs1 and Z1 ). With Vs1 absent, I1 = 0, and by virtue of Eq. (11.39), I2 = 0. Hence, with no voltage drop across Z2 , V2 = Vs2 , and VTh(ab) = V1 =

V2 Vs = 2 . n n

To determine ZTh(ab) , we replace Vs2 with a short circuit and use Eq. (11.40), ZTh(ab) = Zin =

Z2 . n2

(11.41)

(b) Looking to the left from terminals (c, d) involves identically the same process, except that the turns ratio going from terminals (c, d) to terminals (a, b) is now 1/n. The result is diagrammed in Fig. 11-15(c). Example 11-8: Ideal Autotransformer

In an autotransformer, the primary and secondary sides share a part or all of the same winding. Develop expressions for the voltage and current ratios in the autotransformer circuits shown in Fig. 11-16(a) and (b).

“book” — 2015/5/4 — 7:25 — page 619 — #19

11-5 THREE-PHASE TRANSFORMERS

I1

R

Vs

+

+ _

619

Concept Question 11-6: What is the coupling coefficient

of an ideal transformer? (See

N1

V1 N

_

N2

)

I2

+ _

V2

Concept Question 11-7: Are the secondary-to-primary

ZL

voltage and current ratios in an ideal transformer dependent on L1 and L2? (See )

(a) Step-down autotransformer Concept Question 11-8: What is an autotransformer?

I2 R Vs

I1

N1

+ N2 V1 _

+ _

(See

)

+ N V2

Exercise 11-5: Determine the Th´evenin equivalent of the

ZL

circuit to the right of terminals (a, b) in Fig. E11.5.

_

1:4

a

(b) Step-up autotransformer

32 Ω

−j64 Ω + _ 12e j60˚ V

Figure 11-16: Autotransformer circuits.

b Figure E11.5 Solution: (a) For the step-down autotransformer in Fig. 11-16(a),

Answer:

a

V2 N2 N2 = = V1 N N1 + N 2 and

(See

(b) The converse is true for the step-up autotransformer of Fig. 11-16(b). That is,

(step-up autotransformer)

)

Exercise 11-6: An autotransformer is used to step up the voltage by a factor of 10. If N = 200, what are the values of N1 and N2? Answer: N1 = 180 and N2 = 20. (See

V2 N N1 + N2 = = V1 N2 N2 I2 V1 N2 = = . I1 V2 N1 + N 2

3e j60˚ V

b

(step-down autotransformer)

and

_ +

(11.42)

I2 V1 N1 + N2 = = . I1 V2 N2

(2 − j4) Ω

)

(11.43)

11-5 Three-Phase Transformers In Chapter 10 we observed that distribution of power is cheaper (fewer wires) and more efficient (less ohmic loss) when transferred by a three-phase system than by three

“book” — 2015/5/4 — 7:25 — page 620 — #20

620

CHAPTER 11

ILp ILp 1

+

1:n

1

ILs = (1/n) ILp a

Y

+

+

Y

VLp

VLs = nVLp

_

_

1:n

ILs = (1/n) ILp





VLp

+

a

VLs = nVLp

_

c

2

MAGNETICALLY COUPLED CIRCUITS

2

_ c

3 b

3 b

(a) Y-Y transformer

(b) ∆-∆ transformer

1 VLp+ _ 2

Y 3 neutral

:

1 n a



+

b

VLs =

n VLp √3

_

c (c) Y-∆ transformer

Figure 11-17: Three possible connection configurations for three-phase transformers. independent single-phase systems. A similar argument holds for transformers. To step up or step down voltage levels in a three-phase network, it is better to use a single three-phase transformer than three single-phase transformers. Three-phase transformers can be designed to couple any combination of Y- and -configurations on the primary and secondary sides. The three possible combinations are diagrammed in

Fig. 11-17 (Y- and -Y transformers are mirror images of one another). In each case, the line voltages and currents in the secondary circuit are related to those on the primary side. For Y-Y and - connections (Fig. 11-17(a) and (b)), ILp VLs = =n VLp ILs

(Y-Y and - ),

(11.44)

“book” — 2015/5/4 — 7:25 — page 621 — #21

11-5 THREE-PHASE TRANSFORMERS

621

where VLp and VLs are the rms magnitudes of the line voltages on the primary and secondary sides of the transformer, and a similar notation applies to the line currents ILp and ILs . The transformer is assumed to be ideal (lossless and perfectly coupled), in which case conservation of energy requires that STp = STs ,

ILp = nILs = 4 × 77 = 308 A (rms).

Example 11-10: Three-Phase Transformer Circuit

where STp and STs are the total apparent powers at the primary and secondary sides of the transformer. Recall from Eqs. (10.20) and (10.23) that the magnitude of the total apparent power is given by √ ST = 3 VL IL

The line current at the primary side is

Determine Ix in the circuit of Fig. 11-18. Solution: The three mesh equations are −36 + (10 + 50)I1 − 50I2 + V1 = 0,

(Y and ),

(11.45)

for both Y- and -configurations and on either side of the transformer. Consequently, for Y- and -Y transformers (Fig. 11-17(c) and (d)),

ILp VLs n = =√ VLp ILs 3

−50I1 + (30 + 40 + 50)I2 − 30I3 − V2 = 0, and −30I2 + (20 + 30)I3 + V2 − V1 = 0. Additionally, because the transformer dots are on opposite ends [see Fig. 11-14(b) for reference], V2 = −nV1 = −4V1

(Y- ) and

and

(11.46)

√ ILp VLs = = 3n VLp ILs

I2 =

( -Y).

−I1 −I1 = . n 4

After multiple substitutions, we arrive at the solution Ix = I2 = −0.06 A.

Example 11-9: - Transformer

The secondary side of a - three-phase, ideal transformer is connected to a 16 kVA, balanced, three-phase load. If each transformer has a turns ratio of 1 : 4 and the line voltage at the load side is 120 V (rms), determine the line voltage and line current at the primary side.

20 Ω

120 VLs = = 30 V (rms). n 4

Application of the apparent power expressions given by Eq. (11.45) to the secondary side leads to ILs = √

ST 3 VLs

16, 000 =√ = 77 A (rms). 3 × 120

30 Ω

1:4

Solution: For the - connection, Eq. (11.44) gives VLp =

I3

10 Ω

36

0o

V

+ V_1

+ _

Ix

+ V_2

I1

40 Ω I2

50 Ω

Figure 11-18: Circuit of Example 11-10.

“book” — 2015/5/4 — 7:25 — page 622 — #22

622

CHAPTER 11

Concept Question 11-9: What are three-phase trans-

formers used for? (See

MAGNETICALLY COUPLED CIRCUITS

Exercise 11-7: Determine Ix in the circuit of Fig. 11-18 after replacing the 20  resistor with an open circuit.

)

Answer: Ix = −0.097 A. (See

Concept Question 11-10: How is the secondary-to-

)

primary voltage ratio related to the turns ratio n for Y-Y, - , Y- , and -Y configurations? (See )

Summary Concepts • Current flow through a coil in close proximity of a second coil induces a mutual inductance voltage across the second coil through a shared magnetic field. • The dot convention, which accounts for the directions of the windings in the two coupled coils, defines the polarities of the induced mutual-inductance voltages.

• The coupling coefficient of an ideal transformer is k = 1. Its secondary-to-primary voltage and current ratios are defined by the turns ratio n = N2 /N1 . • Three-phase transformers are used to couple any combination of Y- and -configurations on the primary and secondary sides.

• A transformer can be modeled in terms of T- or -equivalent circuits.

Mathematical and Physical Models Magnetic Coupling

i1

R1 υ s1

i2

M

+

+ _

di1 di2 υ1 = L1 dt + M dt _ Change + to − if i1 is CCW

R2

+

L1

L2

Change + to − if i2 is CW

di2 di1 υ2 = L2 dt + M dt _

Change + to − if i2 is CW

+ _

υs2

Change + to − if ii is CCW

(a) Dots on same ends i1

R1 υ s1

+ _

i2

M

+

di1 di2 υ1 = L1 dt − M dt _ Change + to − if i1 is CCW

+

L1

Change − to + if i2 is CW

L2

di2 di1 υ2 = L2 dt − M dt _

Change + to − if i2 is CW

(b) Dots on opposite ends k=√

M L1 L2

R2 + _

Change − to + if ii is CCW

υs2

“book” — 2015/5/4 — 7:25 — page 623 — #23

PROBLEMS

623

Mathematical and Physical Models (continued) Equivalent Circuits

I1 (a) Transformer

Equivalent Inductance L1 a

I2

M

+

+

_

Lx

I1 (b) T-equivalent circuit

L2

L1

V1

V2

_

Ly

+ Lz

V1

_

(c)

Leq = L1 + L2 + 2M

V2

_

Lx = L 1 − M L y = L2 − M Lz = M ∏-equivalent

+

circuit

V1

Lc Lb

La

V2

_

L 1 L2 − M 2 L1 − M L 1 L2 − M 2 Lb = L2 − M L 1 L2 − M 2 Lc = M

Leq = L1 + L2 − 2M Ideal Transformer V2 =n V1

La =

1 I2 = I1 n Zin =

Replace M with −M if dots are on opposite ends.

-equivalent circuit autotransformer coupling coefficient dot convention input impedance linear loosely coupled magnetic flux

L2

M b

I2

+

_

Important Terms

L1

a Leq

I1

L2

M b

I2

+

Leq

V1 1 V2 ZL = 2 = 2 I1 n I2 n

Provide definitions or explain the meaning of the following terms: magnetic flux linkage magnetic permeability magnetically coupled circuit mutual inductance mutual inductance voltage mutual voltage perfectly coupled permeability of free space

primary port primary side reflected impedance secondary port secondary side self inductance step-down transformer step-up transformer

PROBLEMS 10 Ω

Section 11-1: Magnetic Coupling *11.1 For the circuit shown in Fig. P11.1, determine (a) i(t) and (b) the average power absorbed by RL . ∗

+ 30 0 V _ f = 60 Hz

10 mH

T-equivalent circuit three-phase transformer tightly coupled total flux transformer turns ratio two parallel lines voltage transformers

0.2 H

i 1H

RL

Answer(s) available in Appendix G.

Figure P11.1: Circuit for Problem 11.1.

200 Ω

“book” — 2015/5/4 — 7:25 — page 624 — #24

624

CHAPTER 11

MAGNETICALLY COUPLED CIRCUITS

11.2 For the circuit in Fig. P11.2, determine (a) iL (t) and (b) the average power dissipated in RL .

14 Ω

100 μF

6 mH

+ _ 12 cos 377t (V) 10 mH

+ 20 0 V _ f = 60 Hz

30 Ω 30 mH

j6 Ω



j2 Ω

j4 Ω



iL Figure P11.5: Circuit for Problem 11.5.

RL 10 Ω

j2 Ω 4Ω

Figure P11.2: Circuit for Problem 11.2.

11.3

For the circuit in Fig. P11.3, determine Vout .

4Ω 10

0

j3 Ω

+ V _

0

V

+

j1 Ω j6 Ω −j2 Ω



j6 Ω

Vout

j10 Ω



j30 Ω Ix

Figure P11.3: Circuit for Problem 11.3.

60

0

V

+ _

j20 Ω j8 Ω

j5 Ω

20 Ω

Determine Vout in the circuit shown in Fig. P11.4. Figure P11.7: Circuit for Problem 11.7.

11.5 Determine the average power dissipated in the 4  resistor of the circuit in Fig. P11.5. 11.6

Determine Ix in the circuit of Fig. P11.6.

*11.7

Determine Ix in the circuit of Fig. P11.7.



j10 Ω j3 Ω j6 Ω



11.8 Determine the average power dissipated in the 4  resistor of the circuit in Fig. P11.8. *11.9

Ix



_

*11.4

j4 Ω −j10 Ω

Figure P11.6: Circuit for Problem 11.6.

j2 Ω

j4 Ω

30

j6 Ω

+ _

+ 20 0 V _ f = 60 Hz

Determine Vout in the circuit of Fig. P11.9.

j2 Ω

j3 Ω j4 Ω

Figure P11.8: Circuit for Problem 11.8.

8Ω 96

0

+ V _



j1 Ω



+ −j4 Ω

j2 Ω

j6 Ω

−j6 Ω

Figure P11.4: Circuit for Problem 11.4.



Vout

_



“book” — 2015/5/4 — 7:25 — page 625 — #25

PROBLEMS

625

j12 Ω

6Ω 10

0

j6 Ω

j18 Ω

+

+ V _





−j4 Ω



Vout

_

Figure P11.9: Circuit for Problem 11.9.

M

4Ω + 12 cos 2πft (V) _

3 mH

12 mH



Figure P11.10: Circuit for Problem 11.10.

4Ω 100

0

−j2 Ω

j4 Ω



−j4 Ω

+

+ V _

j6 Ω

j12 Ω



Vout

_

Figure P11.11: Circuit for Problem 11.11.

2Ω Vs

+ _

j2 Ω



−j2 Ω

6Ω j6 Ω

j2 Ω

Ix j8 Ω

Figure P11.12: Circuit for Problem 11.12.

11.13 11.10 The circuit shown in Fig. P11.10 uses a 12 V ac source to deliver power to an 8  speaker. If the average power delivered to the speaker is 1.8 W at an audio frequency f = 1 kHz, what is the value of the coupling coefficient k? *11.11 Determine Vout in the circuit in Fig. P11.11. 11.12 Determine Ix in the circuit of Fig. P11.12, given that Vs = 20 30◦ (V).

Determine:

(a) Leq at terminals (a, b) in Fig. P11.13(a). (b) Leq at terminals (a, b) in Fig. P11.13(b). *(c) Leq at terminals (a, b) in Fig. P11.13(c). (d) Leq at terminals (a, b) in Fig. P11.13(d). (e) Leq at terminals (a, b) in Fig. P11.13(e). (f) Leq at terminals (a, b) in Fig. P11.13(f).

“book” — 2015/5/4 — 7:25 — page 626 — #26

626

CHAPTER 11

MAGNETICALLY COUPLED CIRCUITS

5 mH 1 mH

a

5 mH

a Leq

20 mH

4 mH

0.5 mH

Leq

b

6 mH

2 mH

10 mH

b (a)

(b)

a

1H

2H

a 1H 5H

2H

5H

4H

b

b

6H (c)

2H

(d)

a

a

20 mH

20 mH

1H 5H

5H

40 mH

10 mH

b

b (e)

(f) Figure P11.13: Circuits for Problem 11.13.

Sections 11-2 and 11-3: Transformers and Energy

*11.15 Determine (a) the input impedance and (b) the reflected impedance, both at terminals (c, d) in the circuit of Fig. P11.15.

11.14 Determine (a) the input impedance and (b) the reflected impedance, both at terminals (a, b) in the circuit of Fig. P11.14.

a

10 Ω

−j6 Ω

j2 Ω j4 Ω

j2 Ω j6 Ω

b Figure P11.14: Circuit for Problem 11.14.



11.16 For the circuit in Fig. P11.16 (a) determine the Th´evenin equivalent to the left of ZL , (b) choose ZL for maximum power transfer, and (c) compute the average power absorbed by ZL . *11.17 Determine the input impedance Zin of the circuit in Fig. P11.17. 11.18 In the circuit of Fig. P11.18, what should the value of the coupling coefficient k be so that Vout /Vin = 0.49?

“book” — 2015/5/4 — 7:25 — page 627 — #27

PROBLEMS

627

j2 Ω

6Ω j2 Ω

−j2 Ω



12 Ω j6 Ω

c

−j4 Ω d

Figure P11.15: Circuit for Problem 11.15.

j10 Ω 20 Ω

40 Ω j20 Ω

20 Ω j30 Ω

+ _ 120

a

ZL

−j10 Ω 0

V b

Figure P11.16: Circuit for Problem 11.16.

−j2 Ω



2H

j4 Ω 2H

j6 Ω

4H

j8 Ω

Zin

20 Ω

Leq

8H

2H

10 Ω

4H

6H

Figure P11.17: Circuit for Problem 11.17. Figure P11.19: Circuit for Problem 11.19.



Vin

+ _

j1 Ω



+ jωM

j1 Ω −j4 Ω

11.20 For the circuit in Fig. P11.20, determine the complex powers: (a) supplied by the source, (b) stored by the two inductors, and (c) dissipated by the source and load resistors.

Vout

_ Rs Figure P11.18: Circuit for Problem 11.18.

10 *11.19 Apply T- and -transformations to determine Leq in the circuit of Fig. P11.19.

0

+ V _

10 Ω j5 Ω

j4 Ω j15 Ω RL

Figure P11.20: Circuit for Problem 11.20.

10 Ω

“book” — 2015/5/4 — 7:25 — page 628 — #28

628

CHAPTER 11

*11.21 Determine input impedance Zin at terminals (a, b) for the circuit in Fig. P11.21.

a

MAGNETICALLY COUPLED CIRCUITS

11.25 Determine the complex power supplied by the source in the circuit of Fig. P11.25.

1 kΩ

j1 Ω j5 Ω

j5 Ω

24

−j2 Ω

0

−j8 kΩ 1:2

+ V _

16 kΩ

Zin Figure P11.25: Circuit for Problem 11.25.

j1 Ω j5 Ω

12 Ω

j5 Ω

*11.26

b

Determine current Ix in the circuit of Fig. P11.26.

1 kΩ

Figure P11.21: Circuit for Problem 11.21.

Ix 11.22 Determine the average power dissipated in the 10  load in the circuit of Fig. P11.22, given that Vs = 10 0◦ V (rms).



Vs

+ _

j2 Ω

j4 Ω

100

0

−j2 kΩ

4 kΩ

1:2

+ V _

8 kΩ

Figure P11.26: Circuit for Problem 11.26.



+ j6 Ω j1 Ω j5 Ω

10 Ω

Vo

j5 Ω

11.27 Determine currents I1 and I2 in the circuit of Fig. P11.27.

2 kΩ −j2 kΩ

_

I1

4:1

+ 12 0 V _

Figure P11.22: Circuit for Problem 11.22.

Section 11-4: Ideal Transformer 11.23

2 kΩ

j2 kΩ

4 kΩ

I2

+ 24 30 V _

Figure P11.27: Circuit for Problem 11.27.

Determine Vout in the circuit of Fig. P11.23.

1 kΩ + 12 0 V _

*11.28 Determine the average power delivered to RL in the circuit of Fig. P11.28.

1:4

+ 32 kΩ

−j2 kΩ

Vout

_

1 kΩ Figure P11.23: Circuit for Problem 11.23.

120 *11.24 Repeat Problem 11.23 after inserting a 0.5 μF capacitor in series with the 1 k resistor. The angular frequency is 103 rad/s.

0

+ V _

1:4

1 kΩ RL

Figure P11.28: Circuit for Problem 11.28.

2 kΩ

“book” — 2015/5/4 — 7:25 — page 629 — #29

PROBLEMS

629

11.29 Determine the input impedance Zin of the circuit in Fig. P11.29.

20 Ω

1:5

2 kΩ

Section 11-5: Three-Phase Transformers 11.32 A Y- ideal three-phase transformer with a turns ratio of 1 : 10 supplies a 32 kVA load at a line voltage of 208 V. Determine the line voltage and line current at the primary side.

1:4 64 kΩ

Zin

−j32 kΩ

*11.33 A -Y ideal three-phase transformer supplies a 32 kVA load at a line voltage of 240 V. If the line voltage at the primary side is 51.96 V, what is the turns ratio? Potpourri Questions

Figure P11.29: Circuit for Problem 11.29.

11.30 Determine the power absorbed by RL in the circuit of Fig. P11.30.



20

0

1:6

+ V _

RL 144 Ω −j16 Ω

Figure P11.30: Circuit for Problem 11.30.

*11.31 Determine the average power delivered to ZL in the circuit of Fig. P11.31, given that N1 = 50 turns and N2 = 10 turns.

1 kΩ 178

0

+ V _

N1 ZL N2

(10 + j4) kΩ

−j2 Ω

Figure P11.31: Circuit for Problem 11.31.

11.34 What particular features give an MRI advantages over other imaging systems? 11.35 An MRI machine uses a superconducting magnet to create a high magnetic field. For what purpose? Another magnetic field called the RF field is also used. Why?

“book” — 2015/5/4 — 7:26 — page 630 — #1

12

CHAPTER

Circuit Analysis by Laplace Transform

Contents 12-1 12-2 TB28 12-3 12-4 12-5 TB29 12-6 12-7 12-8

Overview, 631 Unit Impulse Function, 631 The Laplace Transform Technique, 633 3-D TV, 637 Properties of the Laplace Transform, 639 Circuit Analysis Procedure, 641 Partial Fraction Expansion, 644 Mapping the Entire World in 3-D, 648 s-Domain Circuit Element Models, 652 s-Domain Circuit Analysis, 655 Multisim Analysis of Circuits Driven by Nontrivial Inputs, 662 Summary, 665 Problems, 665

Objectives Learn to: 

Compute the Laplace transform of a timedependent function.



Perform partial fraction expansion.



Analyze circuits using the Laplace transform technique.

Amps is(t)

6 4

iC(t)

2 0 −2

1

−4 −6 Source and capacitor currents.

2

3

t (s)

“book” — 2015/5/4 — 7:26 — page 631 — #2

12-1

UNIT IMPULSE FUNCTION

631

Overview In Chapters 5 and 6 we examined the transient response to a sudden change in voltage or current in circuits containing RC, RL, and RLC configurations. The excitation sources were dc voltage and current sources, in combination with SPST and SPDT switches. Voltage and current responses in such circuits are characterized by exponential functions of the form e−αt , where α is a damping coefficient that defines the rate at which the response transitions from its initial value immediately after the sudden change to its final value at t = ∞. The timedomain solution method employed in Chapters 5 and 6 is quite satisfactory, so long as the forcing function is a dc source and the differential equation describing the voltages and currents in the circuit is not higher than second order. Now, we introduce the Laplace transform analysis technique, which can be easily applied to a wide range of circuits and any type of realistic forcing function. In fact, the Laplace transform reduces to the phasor transform of Chapter 7 when the circuit sources are time-harmonic sinusoidal functions (and if the Laplacetransform analysis is confined to the steady state component of the overall solution). The phasor-domain technique served us well in Chapters 7–9, but it does not account for the transient component of the circuit response. Because in most ac circuits of interest, the transient component decays rapidly after connecting the sources to the circuit, the steady state solution provided by the phasor-domain technique is all that is needed. However, if we were to seek a solution that incorporates both the transient and steady state components, then the Laplace transform technique is the solution method of choice.

T

0

t

(a) δ(t) and δ(t − T ) δ(t) Area = 1 1/2ε −ε

0

ε (b) Rectangle model

t

u(t) u(t)

1

1 slope = 2ε −ε

12-1 Unit Impulse Function The waveforms commonly encountered in electric circuits include a variety of continuous-time functions—such as the exponential and sinusoid—as well as some discontinuous functions, most notably the step and impulse functions. The step function, defined earlier in Section 5-3, is used to describe mathematically the instantaneous action by a switch to connect or disconnect a source to the circuit. In like manner, the impulse function is a useful mathematical tool for describing a sudden action of very short duration, or for sampling a continuous function at discrete points in time. An example of the latter is when a continuous visual image is stored by recording only 24 images per second, thereby sampling a continuous signal only at specific times. Graphically, the unit impulse function—also known as the delta function δ(t)—is represented by a vertical arrow, as shown in Fig. 12-1(a). If it is located at t = T , it is designated

δ(t − T)

δ(t)

0.5

ε 0 (c) Gradual step model

t

Figure 12-1: Unit impulse function. δ(t − T ). For any fixed value T , the unit impulse function is defined through the combination of two properties: δ(t − T ) = 0

for t = T ,

(12.1a)

∞ δ(t − T ) dt = 1.

(12.1b)

−∞

 The first property states that the unit impulse function δ(t − T ) is zero everywhere, except at its own location (t = T ), but its value is infinite at that location! The second property states that the total area under the unit impulse function is equal to 1, regardless of its location. 

“book” — 2015/5/4 — 7:26 — page 632 — #3

632

CHAPTER 12

To visualize the meaning of the second property, we can represent the unit impulse function by the graphical rectangle shown in Fig. 12-1(b), with the understanding that δ(t) is defined in the limit as ε → 0. The rectangle is of width w = 2ε and height h = 21 ε, so its area is always unity, even as ε → 0. Since δ(t − T ) in the integral of Eq. (12.1b) is, by definition, equal to zero everywhere except over an infinitesimally narrow range surrounding t = T , Eq. (12.1b) can be reexpressed as T+ε

δ(t − T ) dt = 1.

(12.2)

T −ε

12-1.1

12-1.2

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Sampling Property of δ(t)

As was noted earlier, multiplying an impulse function by a constant A gives a scaled impulse of area A. Now we consider what happens when a continuous-time function x(t) is multiplied by δ(t). Since δ(t) is zero everywhere except at t = 0, it follows that x(t) δ(t) = x(0) δ(t),

provided that x(t) is continuous at t = 0. By extension, multiplication of x(t) by the time-shifted impulse function δ(t − T ) gives

Relationship Between u(t) and δ(t)

x(t) δ(t − T ) = x(T ) δ(t − T ).

According to the rectangle model displayed in Fig. 12-1(b), δ(t) = 1/(2ε) over the narrow range −ε < t < ε. For the gradual step model of u(t) shown in Fig. 12-1(c), its slope also is 1/(2ε). Hence, du(t) = δ(t). dt

(12.3)

Even though this relationship between the unit impulse and unit step functions was obtained on the basis of specific geometrical models for δ(t) and u(t), its validity can be demonstrated to be true always. The corresponding expression for u(t) is

δ(τ ) dτ,

One of the most useful features of the impulse function is its sampling property. For any function x(t) known to be continuous at t = T : ∞ x(t) δ(t − T ) dt = x(T ).

(12.4)

(sampling property)

where τ is a dummy integration variable introduced so as to avoid confusion between the integration variable τ and the upper limit of the integral t in Eq. (12.4). For the time-shifted case,

Derivation of the sampling property relies on Eqs. (12.1b) and (12.8): ∞

∞ x(t) δ(t − T ) dt =

t δ(τ − T ) dτ.

(12.5a)

−∞

d [u(t − T )] = δ(t − T ), dt

∞ = x(T )

t (12.6)

δ(t − T ) dt = x(T ).

−∞

(12.5b)

By extension, a scaled impulse A δ(t) has an area A and

A δ(τ ) dτ = A u(t).

x(T ) δ(t − T ) dt −∞

−∞

−∞

(12.9)

−∞

−∞

u(t − T ) =

(12.8)

 Multiplication of a time-continuous function x(t) by an impulse located at t = T generates a scaled impulse of magnitude x(T ) at t = T , provided x(t) is continuous at t = T . 

t u(t) =

(12.7)

Concept Question 12-1: How is u(t) related to δ(t)?

(See

)

Concept Question 12-2: Why is Eq. (12.9) called the sampling property of the impulse function? (See )

“book” — 2015/5/4 — 7:26 — page 633 — #4

12-2 THE LAPLACE TRANSFORM TECHNIQUE

633

Exercise 12-1: If x(t) is the rectangular pulse shown in

Fig. E12.1(a), determine its time derivative x  (t) and plot it.

x(t) 2

3

t (s)

4

(a) x(t) x′(t) 2 δ(t − 3) t (s) −2 δ(t − 4) (b) x′(t) Figure E12.1 Answer: x(t) = 2δ(t − 3) − 2δ(t − 4). (See

)

12-2 The Laplace Transform Technique  A domain transformation is a mathematical process that converts a set of variables from their domain into a corresponding set of variables defined in another domain.  In reality, a circuit functions entirely in the time domain, with both its inputs (sources) and outputs (voltages and currents) expressed as functions of the time variable t. In the top horizontal sequence depicted in Fig. 12-2, application of KCL and KVL leads to the differential equation characterizing the output of interest, and its solution then yields the desired response. All mathematical steps are performed entirely in the time domain. A transform is a mathematical operator that converts functions, such as υ(t) and i(t), into counterpart functions defined in another domain. The Laplace transform is one such operator; it converts a function υ(t) defined in the time domain into a counterpart V(s) defined in another domain called the s-domain. The two domains may be thought of as “parallel universes” and a transformation is a “transport” between the two universes. By applying the Laplace transform, a circuit (or its associated differential equation) can be transformed to the s-domain. The transformed circuit is characterized by an algebraic equation—instead of a differential equation— with a relatively straightforward solution. By transforming

Time Domain Time-Domain Circuit

Differential Equation

Inverse Laplace transform

Laplace transform s-Domain Circuit

Time-Domain Solution

Algebraic Equation

s-Domain Solution

s-Domain Figure 12-2: The top horizontal sequence involves solving a differential equation entirely in the time domain. The bottom horizontal sequence involves a much easier solution of a linear equation in the s-domain.

“book” — 2015/5/4 — 7:26 — page 634 — #5

634

CHAPTER 12

the solution back to the time domain, we end up with the same solution that we would have obtained had we solved the differential equation directly in the time domain. Even though the s-domain route involves transforming the circuit to the sdomain and transforming the solution to the time domain (which is called an inverse Laplace transformation), the overall solution process is considerably easier to implement than the traditional differential-equation solution method.

Solution Procedure: Laplace Transform

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

f (t) defined in the time domain into a function F(s) defined in the s-domain. Functions f (t) and F(s) are called a Laplace transform pair. The uniqueness property of the Laplace transform states:  A given f (t) has a unique F(s), and vice versa.  The uniqueness property can be expressed in symbolic form by f (t)

Step 1: The circuit is transformed to the Laplace domain—also known as the s-domain. Step 2: In the s-domain, application of KVL and KCL yields a set of algebraic equations. Step 3: The equations are solved for the variable of interest. Step 4: The s-domain solution is transformed back to the time domain. After introducing the Laplace transform and exploring its properties, we will demonstrate its capabilities by applying the outlined four-step procedure to analyze several types of passive and active circuits.

12-2.1

Definition of the Laplace Transform

The symbol L [f (t)] is a short-hand notation for “the Laplace transform of function f (t).” Usually denoted F(s), the Laplace transform is defined by ∞ F(s) = L [f (t)] =

f (t) e−st dt,

(12.12a)

The two-way arrow is a short-hand notation for the combination of the two statements L −1 [F(s)] = f (t).

L [f (t)] = F(s),

(12.12b)

The first statement asserts that F(s) is the Laplace transform of f (t), and the second one asserts that the inverse Laplace L−1 [ ]) of F(s) is f (t). transform (L Because the lower limit on the integral in Eq. (12.10) is 0− , F(s) is called a one-sided transform, in contrast with the twosided transform for which the lower limit is −∞. When we apply the Laplace transform technique to electric circuits, we select the start time for the circuit operation as t = 0, so the single-sided transform is plenty suitable for our intended use, and we will adhere to it exclusively in this book. Moreover, unless noted to the contrary, it will be assumed that f (t) is always multiplied by an implicit invisible step function u(t). The inquisitive reader may ask why we use 0− , instead of simply 0, as our lower limit in the integral of Eq. (12.10). We use it as a reminder that the integration can include initial conditions at t = 0− , which may be associated with the voltage across a capacitor or the current through an inductor.

(12.10)

12-2.2

0−

where s is a complex variable with a real part σ and an imaginary part ω: s = σ + j ω.

F(s).

(12.11)

Given that the exponent st has to be dimensionless, s has the unit of inverse second, which is the same as Hz or rad/s. Moreover, since s is a complex quantity, it is often termed complex frequency. In view of the definite limits on the integral in Eq. (12.10), the outcome of the integration will be an expression that depends on a single variable, s. The transform operation converts a function

Convergence Condition

Depending on the functional form of f (t), the Laplace transform integral given by Eq. (12.10) may or may not converge to a finite value. If it does not, the Laplace transform does not exist. Convergence requires that ∞ |f (t) e

−st

∞ | dt =

0−

|f (t)||e−σ t ||e−j ωt | dt

0−

∞ = 0−

|f (t)|e−σ t dt < ∞,

(12.13)

“book” — 2015/5/4 — 7:26 — page 635 — #6

12-2 THE LAPLACE TRANSFORM TECHNIQUE

635

for some real value of σ . We used the fact that |e−j ωt | = 1 for any value of ωt and, since σ is real, |e−σ t | = e−σ t . If σc is the smallest value of σ for which the integral converges, then the region of convergence is σ > σc . Fortunately, this convergence issue is somewhat esoteric to circuit analysts and designers because the waveforms of the excitation sources used in electric circuits are such that they do satisfy the convergence condition for all values of σ , and hence, their Laplace transforms do exist.

12-2.3

f1(t) Step A

Inverse Laplace Transform

f2(t)

Equation (12.10) allows us to obtain Laplace transform F(s) corresponding to time function f (t). The inverse process, denoted L −1 [F(s)], allows us to perform an integration on F(s) to obtain f (t):

f (t) = L −1 [F(s)] =

1 2πj

A Rectangle

T1

σ +j ∞

F(s) est ds,

t

T (a)

t

(b)

(12.14)

σ −j ∞

where σ > σc . The integration, which has to be performed in the two-dimensional complex plane, is rather cumbersome and to be avoided if an alternative approach is available for converting F(s) into f (t). Fortunately, there is an alternative approach. Instead of applying Eq. (12.14), we can generate a table of Laplace transform pairs for all of the time functions commonly encountered in electric circuits, and then use it, sort of like a look-up table, to transform the s-domain solution to the time domain. The validity of this approach is supported by the uniqueness property of the Laplace transform, which guarantees a one-to-one correspondence between every f (t) and its corresponding F(s). The details of the inverse-transform process are covered in Section 12-4.

T2

f3(t) A Impulse T

t

(c) Figure 12-3: Singularity functions for Example 12-1.

Application of Eq. (12.10) gives Example 12-1: Laplace Transforms of Singularity Functions

The step, rectangle, and impulse waveforms displayed in Fig. 12-3 are known as singularity functions, because either they or their time derivatives exhibit discontinuities. Determine their Laplace transforms. Solution: (a) The step function in Fig. 12-3(a) is given by f1 (t) = A u(t − T ).

∞ F1 (s) =

f1 (t) e−st dt

0−

∞ =

A u(t − T ) e−st dt

0−

∞ e

=A T

−st

 A −st ∞ A −sT dt = − e  = . e s s T

For the special case where A = 1 and T = 0 (the step occurs

“book” — 2015/5/4 — 7:26 — page 636 — #7

636

CHAPTER 12

at t = 0), the transform pair becomes 1 . s

u(t)

(12.15)

(b) The rectangle function in Fig. 12-3(b) can be constructed as the sum of two step functions:

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Solution: Inclusion of u(t) in the expression for f (t) is simply a reminder of the implicit assumption common to all excitations considered in this chapter, namely that f (t) = 0 for t < 0− . The solution is facilitated by expressing cos ωt in terms of complex exponentials (see Table 7-2), namely cos ωt =

f2 (t) = A[u(t − T1 ) − u(t − T2 )], and its Laplace transform is ∞ F2 (s) =

Use of this expression in Eq. (12.10) gives

A[u(t − T1 ) − u(t − T2 )]e−st dt

∞ F(s) =

0−

∞ =A

u(t − T1 ) e

−st

∞ dt − A

0−

=

1 j ωt [e + e−j ωt ]. 2

cos ωt u(t) e−st dt

0−

u(t − T2 ) e

−st

⎡∞ ⎤  ∞ 1 = ⎣ ej ωt e−st dt + e−j ωt e−st dt ⎦ 2

dt

0−

A −sT1 − e−sT2 ]. [e s

1 = 2

(c) The impulse function in Fig. 12-3(c) is given by f3 (t) = A δ(t − T ),



0

0

e(j ω−s)t e−(j ω+s)t + jω − s −(j ω + s)

∞  s  .  = 2  s + ω2 0

Hence,

and the corresponding Laplace transform is ∞ F3 (s) =

[cos(ωt)] u(t)

A δ(t − T ) e−st dt

s . s2 + ω 2

(12.17)

0− T+ε

δ(t − T ) e−st dt

=A

Concept Question 12-3: Is the uniqueness property of

T −ε

= Ae−sT

the Laplace transform unidirectional or bidirectional? Why is that significant? (See )

T+ε

δ(t − T ) dt = Ae−sT ,

T −ε

Concept Question 12-4: Is convergence of the Laplace

where we have used the procedure introduced earlier in connection with Eq. (12.9). For the special case where A = 1 and T = 0, the Laplace transform pair simplifies to δ(t)

1.

(12.16)

Example 12-2: Laplace Transform of cos ωt

Obtain the Laplace transform of f (t) = [cos(ωt)] u(t).

transform integral in doubt when applied to circuit analysis? If not, why not? (See )

Exercise 12-2: Determine the Laplace transform of (a) [sin(ωt)] u(t), (b) e−at u(t), and (c) t u(t). Assume all waveforms are zero for t < 0. ω Answer: (a) [sin(ωt)] u(t) , s2 + ω 2 1 1 (b) e−at u(t) , (c) t u(t) . (See ) s+a s2

“book” — 2015/5/4 — 7:26 — page 637 — #8

TECHNOLOGY BRIEF 28: 3-D TV

637

Technology Brief 28 3-D TV Attempts to produce stereoscopic perception of depth in images have been recorded since at least the mid-19th century. Stereopsis is the impression of depth that arises when humans and other animals view the world using two eyes. Since each eye is at a slightly different location with respect to any object in a viewed scene, the brain can use the difference between the left and right eye images to extract information about depth (and, thus, perceived three dimensionality). Stereopsis was first described in detail by Charles Wheatstone in the 1830s (although it had been observed but not properly understood during the Italian Renaissance). A variety of devices, usually termed stereoscopes, were constructed during the Victorian era which presented viewers with slightly different images to the left and right eyes (usually by projecting them through lenses into each eye separately).

Cyan pixels (or projection)

Magenta filter (a) Anaglyphic 3-D

Pixels with polarizing filter Leftpolarized filter

Rightpolarized filter

(b) Polarizing (passive) 3-D

Left-eye image

Right-eye image

Pixels

(c) Active shutter 3-D

Slits Opaque film Pixels (d) Parallax barrier 3-D

The New Rise of 3-D TV The rapid maturation of flat-screen television, leading to very high resolution, very fast refresh times, high contrast ratios and deep color reproduction, enabled a new resurgence in 3-D technologies in the last few years. Several technologies are currently competing for market dominance (with others in development).

Glasses

Cyan filter

Anaglyphic 3-D The rise of modern cinema saw several additional attempts to convey stereoscopic information to the big screen.The most popular of these, at its peak in the 1950s and 60s, was anaglyphic projection (Fig. TF28-1(a)). Viewers watching a projected movie (or television screen) wore glasses with different color filters in front of each eye (usually cyan and red). Two images were projected simultaneously on the screen, such that only one image could pass each filter and be perceived by the eye. Since each of the two images or films had been captured by cameras slightly offset from each other (to mimic the separation of human eyes), the brain stitched these images together somewhat naturally and perceived depth and “3-D” in the film. Anaglyphic technologies suffer somewhat in that they do not provide perfect color reconstruction and can often produce blurry or ghost images (depending on the quality of the filters used).

Magenta pixels (or projection)

Lens array Pixels (e) Lenticular-lens array 3-D Figure TF28-1: Various techniques for realizing 3-D imaging.

“book” — 2015/5/4 — 7:26 — page 638 — #9

638

Polarizing 3-D So-called polarizing or passive 3-D TV is in some ways similar to anaglyphic systems. Instead of employing color filters, 3-D passive TV glasses use light polarization to deliver different images to the left and right eye (Fig. TF28-1(b)). Light is, of course, electromagnetic radiation perceived by the human eye. Traveling waves of electromagentic radiation are composed of oscillating electric and magnetic fields with specific orientations or polarizations (this does not refer to the direction of travel of the light, but rather, the directions in which the light’s electric and magnetic fields oscillate as the wave travels through space). The details of light polarization can be complex, but of importance to us is the fact that polarizations can be complementary; for example, light can travel with clockwise polarity or counter-clockwise polarity with respect to its source. The human eye cannot tell the difference between different polarizations of light. Passive 3-D television sets contain polarizing filters placed in front of the pixels in the display; with these filters, half of the pixels in the television emit clockwise polarized light and the other half produce counterclockwise polarized light. The viewer wears glasses which let only one polarity of light through to the left eye and the other to the right eye. In this way, each eye is presented with a different image. This type of 3-D TV has advantages in that the glasses are very lightweight. Historically, the main drawback of 3-D images was the decrease in resolution of the image (since each eye received half of the total pixels of the television). Recent advances, however, may be solving this problem. One approach, made possible by the speed of modern pixels, is to present each eye with half of each full-resolution stereoscopic image (as above) and then, very quickly, present the eyes with the other half of a full resolution image. This requires televisions that can refresh images at 120 Hz (120 times each second or a new image every 8.3 milliseconds); this is about double the speed at which a human eye can perceive flicker!

Active Shutter 3-D Active shutter 3-D or alternate frame-sequencing 3-D sets also use glasses but they tend to be heavier and more expensive (Fig. TF28-1(c)). This type of technology uses a normal flat-screen TV (but it must be fast) to display the images intended for the left eye alternating in time with the images intended for the right eye. In other words, while watching a movie, the television displays a frame intended for the left eye followed by a frame intended for

TECHNOLOGY BRIEF 28: 3-D TV the right eye, and so on. The glasses hold LCD screens over each eye and receive a synchronization signal from the television (either infrared or radio frequency). In sync with the TV, the glasses block light to the left eye (by darkening the LCD), then block light to the right eye, and so on, repeating this sequence 24 times a second or faster. In this way, each eye receives the stereoscopic pair intended for it at full resolution. Unlike traditional passive 3-D TVs, all pixels are used for each frame of the image.

Parallax Barrier 3-D Parallax-barrier displays are a glasses-free 3-D display technology that has been around for a number of years but is beginning to make it into prototype flat-panel televisions. Parallax-barrier technology was used in the Nintendo 3-DS hand-held and several 3-D smartphones. The idea behind parallax barrier technology is shown in Fig.TF28-1(d). An opaque film with precisely aligned slots is fabricated over the display pixels; the slots are intended to block light from some pixels from reaching the left eye and to block light from other pixels from reaching the right eye for a viewer standing directly in front of the display. The principal advantage of parallax-barrier displays is that no glasses are needed; anyone standing in front of the TV sees images in stereoscopic 3-D. The two principal disadvantages are the halved resolution (similar to traditional passive 3-D, as explained above), since light from only about half the pixels reach each eye, and the limited viewing angle for 3-D.Typical prototypes only work within a 20◦ angle on either side of center, making it less attractive for consumer use (although the technology is evolving fast).

Lenticular-Lens Arrays Lenticular-lens arrays (Fig. TF28-1(e)) work in a similar manner to parallax viewing except that the light from a given pixel is focused onto the right or left eyes (as opposed to blocked) by an array of precisely fabricated lenses (a very thin plastic sheet is usually molded into a regular array of lenses) sitting over the display. Lenticularlens displays currently suffer from similar drawbacks to parallax-barrier displays and are currently very expensive. As with parallax-barrier technology, several companies are actively pursuing this technology and prices may drop rapidly as the technology matures.

“book” — 2015/5/4 — 7:26 — page 639 — #10

12-3

PROPERTIES OF THE LAPLACE TRANSFORM

12-3 Properties of the Laplace Transform

639

12-3.2 Time Shift If t is shifted by T along the time axis, with T ≥ 0, then

The Laplace transform has a number of useful, universal properties that apply to any function f (t), greatly facilitating the process of transforming a circuit from the t-domain to the s-domain. This section will conclude with a table outlining several universal properties of the Laplace transform, to which we will be making frequent reference throughout this chapter. Some of these properties are intuitively obvious, while others may require some elaboration.

T ≥ 0.

The validity of this property is demonstrated as follows: ∞

∞

f (t)

F(s),

(12.18)

=

then the transform of the time-scaled function f (at) is 1 s f (at) F , a > 0. a a (time-scaling property)

∞ =

(12.19)

L [f (at)] =

f (at) e−st dt.

(12.20)

0−

In the integral, if we set τ = at and dt =

1 a

dτ , we have

f (τ ) e−(s/a)τ dτ

0−

∞ 0−

f (τ ) e−μ τ dτ,

with μ =

f (x) e−s(x+T ) dx

0

To prove Eq. (12.19), we start with the standard definition of the Laplace transform given by Eq. (12.10): ∞

f (t − T ) e−st dt

T

 The time scaling property states that stretching the time axis by a factor a corresponds to shrinking the s axis and the amplitude of F(s) by the same factor, and vice versa. 

1 = a

f (t − T ) u(t − T ) e−st dt

0−

If

1 L [f (at)] = a

(12.23)

(time-shift property)

L [f (t − T ) u(t − T )] =

12-3.1 Time Scaling

∞

e−T s F(s),

f (t − T ) u(t − T )

s . (12.21) a

The definite integral is identical in form to the Laplace transform definition given by Eq. (12.10), except that the dummy variable is τ instead of t, and the coefficient of the exponent is μ = s/a instead of just s. Hence, 1 1 s L [f (at)] = F(μ) = F , a > 0, (12.22) a a a which proves the time-scaling property defined by Eq. (12.19).

=e

−T s

=e

−T s

∞

f (x) e−sx dx

0

F(s),

(12.24)

where we made the substitutions t − T = x and dt = dx, and then applied the definition for F(s) given by Eq. (12.10). To illustrate the utility of the timeshift property, we consider the cosine function of Example 12-2, where it was shown that s [cos(ωt)] u(t) . (12.25) s2 + ω 2 According to Eq. (12.23), e−T s

[cos ω(t − T )] u(t − T )

s . s2 + ω 2

(12.26)

Had we analyzed a linear circuit driven by a sinusoidal voltage source that started at t = 0, and then we wanted to reanalyze it anew, but wanted to delay both the cosine function and the start time by T , Eq. (12.26) would provide an expedient solution to obtaining the transform of the delayed cosine function. Exercise 12-3: Determine L [sin ω(t − T ) u(t − T )] for

T ≥ 0.

Answer:

e−T s

(See

)

s2

ω . + ω2

“book” — 2015/5/4 — 7:26 — page 640 — #11

640

12-3.3

CHAPTER 12

Frequency Shift

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

To verify the validity of Eq. (12.28), we start with the standard definition for the Laplace transform:

According to the time-shift property, if t is replaced with (t − T ) in the time domain, F(s) gets multiplied by e−T s in the s-domain. Within a (−) sign, the converse is also true: if s is replaced with (s + a) in the s-domain, f (t), gets multiplied by e−at in the time domain. Thus,

∞



L [f ] = 0−

df −st dt. e dt

(12.29)

Integration by parts, with e−at f (t)

F(s + a).

dx = −se−st dt, x = e−st , df dy = dt, and y = f, dt

(12.27)

(frequency shift property)

gives

Proof of Eq. (12.27) is part of Exercise 12-4. Concept Question 12-5: According to the time scaling

property of the Laplace transform, “stretching the time axis corresponds to shrinking the s axis.” What does that mean? (See )

∞ L[f  ] = xy 0− −

y dx 0−

=e

−st

∞ f (t)0− −

∞

−s f (t) e−st dt

0−

Concept Question 12-6: Explain the similarities and



= −f (0 ) + s F(s),

differences between the time-shift and frequency-shift properties of the Laplace transform. (See )

Exercise 12-4: (a) Prove Eq. (12.27) and (b) apply it to

∞

(12.30)

which is equivalent to Eq. (12.28). Higher derivatives can be obtained by repeating the application of Eq. (12.28). For the second derivative of f (t),

determine L {[e−at cos(ωt)] u(t)}.

Answer: (a) (See

C3

),

(b) [e−at cos(ωt)] u(t)

s+a . (See (s + a)2 + ω2

f  =

)

d 2f s2 F(s) − s f (0− ) − f  (0− ), dt 2 (second-derivative property) (12.31)

where f  (0− ) is the derivative of f (t), evaluated at t = 0− .

12-3.4 Time Differentiation 12-3.5 Time Integration Differentiating f (t) in the time domain is equivalent to: (a) multiplying F(s) by s in the s-domain, and then (b) subtracting f (0− ) from s F(s):

df f = s F(s) − f (0− ). dt (time-differentiation property)

Integration of f (t) in the time domain is equivalent to dividing F(s) by s in the s-domain: t f (τ ) dτ

(12.28) 0−

1 F(s). s

(time-integration property)

(12.32)

“book” — 2015/5/4 — 7:26 — page 641 — #12

12-4

CIRCUIT ANALYSIS PROCEDURE

641 Since u(t) is equal to the time integral of δ(t), and r(t) is the time integral of u(t), it follows that

Application of the Laplace transform definition gives ⎡ L⎣

t

⎡ ⎤ ∞  t f (τ ) dτ ⎦ = ⎣ f (τ ) dτ ⎦ e−st dt, ⎤

0−

0−

t

(12.33) u(t) =

0−

δ(τ ) dτ

1 , s

u(τ ) dτ

1 . s2

0−

where, for the sake of clarity, we changed the dummy variable in the inner integral from t to τ . Integration by parts with

and t

t x=

f (τ ) dτ

r(t) =

dx = f (τ ) dτ,

0−

0

dy = e−st dt

y=−

e−st , s

Table 12-1 provides a summary of the major properties of the Laplace transform, and Table 12-2 provides a list of Laplace transforms of commonly encountered time functions.

leads to ⎡ L⎣

t



Exercise 12-5: Obtain the Laplace transform of (a) f1 (t) = 2(2 − e−t ) u(t) and (b) f2 (t) = e−3t cos(2t + 30◦ ) u(t).

f (τ ) dτ ⎦

0−

∞ = xy 0− − ⎡

e−st = ⎣− s

y dx 0−

t 0−

2s + 4 , s(s + 1) 0.866s + 3.6 (b) F2 (s) = 2 . (See s + 6s + 13 Answer: (a) F1 (s) =

∞ ⎤∞  ∞  1 1  ⎦ f (τ ) dτ  + f (t) e−st dt = F(s). s s  − 0−

0

12-4

)

Circuit Analysis Procedure

(12.34) Both limits on the first term on the right-hand side yield zero values; at the upper limit,  e−st t=∞ = 0, and at the lower limit,

Now that we have learned how to transform a time-domain function f (t) to its Laplace counterpart F(s), we shall demonstrate the basic steps of the Laplace transform technique by analyzing a relatively simple circuit. Figure 12-4 contains a series RLC circuit, with no stored energy, connected to a dc voltage source Vo via a SPST switch that closes at t = 0. Hence, the source should be represented as υs (t) = Vo u(t).

0−

(12.35)

f (τ ) dτ = 0. Fundamentally, the Laplace transfer technique consists of four steps:

0−

To illustrate the utility of the time-integration property given by Eq. (12.32), we consider the relationships between δ(t), u(t), and r(t). From Eq. (12.16), we have δ(t)

1.

Step 1: Apply KCL and/or KVL to obtain the integrodifferential equation(s) of the circuit for t ≥ 0 For the circuit in Fig. 12-4(a), KVL at t ≥ 0 gives R i(t) + υC (t) + L

di = Vo u(t), dt

(12.36)

“book” — 2015/5/4 — 7:26 — page 642 — #13

642

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Table 12-1: Properties of the Laplace transform (f (t) = 0 for t < 0− ). F(s) = L[f (t)]

f (t)

Property

1. Multiplication by constant K f (t) 2. Linearity

K F(s)

K1 f1 (t) + K2 f2 (t) f (at),

3. Time scaling 4. Time shift

K1 F1 (s) + K2 F2 (s) 1 s F a a

a>0

e−T s F(s),

f (t − T ) u(t − T )

5. Frequency shift

e−at f (t)

6. Time 1st derivative

f = f  =

7. Time 2nd derivative

F(s + a)

df dt

s F(s) − f (0− )

d2f dt 2

s2 F(s) − sf (0− ) − f  (0− )

t

1 F(s) s

f (τ ) dτ

8. Time integral 0−

9. Frequency derivative

t f (t)

10. Frequency integral

f (t) t

where i(t) is the current flowing through the loop and υC (t) is the voltage across C. By invoking the i-υ relationship for C, Eq. (12.36) becomes

T ≥0

d F(s) − ds ∞ F(s ) ds s

Tables 12-1 and 12-2, as follows: R i(t)

R I(s) (multiplication by constant),

⎡ 1 Ri + ⎣ C

t 0−

⎤ di i dt + υC (0 )⎦ + L = Vo u(t), dt −

(12.37)

which now contains a single dependent variable, i(t). Step 2: Define Laplace transform currents and voltages corresponding to the time-domain currents and voltages and then transform the equation to the s-domain We designate I(s) as the s-domain counterpart of i(t), i(t)

I(s).

(12.38)

To transform Eq. (12.37) to the s-domain, we apply the appropriate property or Laplace transformation (LT) from

1 C

t i dt

1 I(s) C s

(time-integral property),

υC (0− )

υC (0− ) s

(LT of a constant),

0−

L

di dt

L[s I(s) − i(0− )] (time derivative property),

Vo (LT of a constant). s The opening paragraph of this section stated that the circuit had no stored energy prior to t = 0. Hence, υC (0− ) = 0 and i(0− ) = 0. Replacing each of the terms in Eq. (12.37) with its s-domain counterpart leads to Vo u(t)

RI +

Vo I + LsI = Cs s

(s-domain).

(12.39)

“book” — 2015/5/4 — 7:26 — page 643 — #14

12-4

CIRCUIT ANALYSIS PROCEDURE

643

Table 12-2: Examples of Laplace transform pairs for T ≥ 0. Note that multiplication by u(t) guarantees that f (t) = 0 for t < 0− . Laplace Transform Pairs f (t)

F(s) = L[f (t)]

1

δ(t)

1

1a

δ(t − T )

e−T s

2

1 or u(t)

2a

u(t − T )

3

e−at u(t)

3a

e−a(t−T ) u(t − T )

4

t u(t)

4a

(t − T ) u(t − T )

5

t 2 u(t)

6

te−at u(t)

7

t 2 e−at u(t)

8

t n−1 e−at u(t)

9

sin ωt u(t)

10

sin(ωt + θ ) u(t)

11

cos ωt u(t)

12

cos(ωt + θ ) u(t)

13

e−at sin ωt u(t)

14

e−at cos ωt u(t)

1 s e−T s s 1 s+a e−T s s+a 1 s2 e−T s s2 2 s3 1 (s + a)2 2 (s + a)3 (n − 1)! (s + a)n ω 2 s + ω2 s sin θ + ω cos θ s2 + ω 2 s s2 + ω 2 s cos θ − ω sin θ s2 + ω 2 ω (s + a)2 + ω2 s+a (s + a)2 + ω2

15

2e−at cos(bt − θ ) u(t)

2t n−1 −at cos(bt − θ ) u(t) e (n − 1)! Note: (n − 1)! = (n − 1)(n − 2) . . . 1. 16

e−j θ ej θ + s + a + jb s + a − jb ej θ e−j θ + (s + a + j b)n (s + a − j b)n

“book” — 2015/5/4 — 7:26 — page 644 — #15

644

CHAPTER 12

R

Hence,

i

i(t) = 4te−5t u(t),



t=0 + _ V0 = 1.6 V

C

0.1 F

0.4 H υs(t) = V0 u(t)

L

(12.41)

and its plot is displayed in Fig. 12-4(b). In this particular example, the expression for I(s) given by Eq. (12.40) matches one of the entries available in Table 12-2, but what should we do if it does not? We have two options: (1) we can apply the inverse Laplace transform relation given by Eq. (12.14), which in general involves a rather cumbersome integration, or

(a) RLC circuit

(2) we can apply the partial-fraction-expansion method to rearrange the expression for I(s) into a sum of terms, each of which has an appropriate match in Table 12-2. This latter approach is the subject of the next section.

i (A) 0.3 0.25 i(t) = 4te−5t u(t)

0.2

12-5

0.15 0.1 0.05 0 0

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

0.2

0.4

0.6

0.8

1

t (s)

(b) i(t) Figure 12-4: The dc source, in combination with the switch, constitutes an input excitation υs (t) = Vo u(t).

Partial Fraction Expansion

Let us assume that after transforming the integrodifferential equation associated with a circuit of interest to the s-domain, and then solving it for the voltage or current whose behavior we wish to examine, we end up with an expression F(s). Our next step is to inverse transform F(s) to the time domain, thereby completing our solution. The degree of mathematical difficulty associated with the implementation of the inverse transformation depends on the mathematical form of F(s). Consider, for example, the expression F(s) =

4 6 8 + + 2 . s + 2 (s + 5)2 s + 4s + 5

(12.42)

The inverse transform, f (t), is given by Step 3: Solve for the variable of interest in the s-domain Solving for I(s), and then replacing R, L, C, and Vo with their numerical values, leads to I(s) =

=

Vo   1 R L s2 + s + L LC s2

4 4 = . + 10s + 25 (s + 5)2

By comparison with the entries in Table 12-2, we note that: (12.40)

Step 4: Transform the solution back to the time domain with the help of Tables 12-1 and 12-2 According to entry #6 in Table 12-2, L −1



 1 = te−at u(t). (s + a)2

f (t) = L −1 [F(s)]       4 6 8 −1 −1 −1 =L +L +L . s+2 (s + 5)2 s2 + 4s + 5 (12.43)

(a) The first term in Eq. (12.43), 4/(s + 2), is functionally the same as entry #3 in Table 12-2, with a = 2. Hence,   4 −1 L (12.44a) = 4e−2t u(t). s+2 (b) The second term, 6/(s + 5)2 , is functionally the same as entry #6 in Table 12-2, with a = 5. Thus,   6 L −1 = 6te−5t u(t). (12.44b) (s + 5)2

“book” — 2015/5/4 — 7:26 — page 645 — #16

12-5

PARTIAL FRACTION EXPANSION

645

(c) The third term, 1/(s2 + 4s + 5), is similar but not identical in form to entry #13 in Table 12-2. However, it can be rearranged to assume the proper form: s2

1 1 = . + 4s + 5 (s + 2)2 + 1

Consequently,   8 = 8e−2t sin t u(t). L −1 (s + 2)2 + 1

(12.44c)

Combining the results represented by Eqs. (12.44a–c) gives: f (t) = [4e−2t + 6te−5t + 8e−2t sin t] u(t).

(12.45)

The preceding example demonstrated that the implementation of the inverse Laplace transform is a rather painless process, so long as the expression for F(s) is composed of a series of terms similar to those in Eq. (12.42). If, however, F(s) is not in the proper form, we will need to reconfigure it before we can apply the inverse transform. Specifically, F(s) should be expanded into a sum of partial fractions by applying the applicable recipe from among those outlined in the forthcoming subsections.

12-5.1

Distinct Real Poles

Consider the s-domain function F(s) =

s2 − 4s + 3 . s(s + 1)(s + 3)

A1 A2 A3 + + , s (s + 1) (s + 3)

(12.47)

where A1 to A3 are expansion coefficients to be determined shortly. Equating the two functional forms of F(s), we have A1 A2 A3 s2 − 4s + 3 + + = . s (s + 1) (s + 3) s(s + 1)(s + 3)

After reduction, the expression becomes   A1 (s + 1) A3 (s + 1)  + A2 + s (s + 3) s=−1  2  (s − 4s + 3)  = .  s(s + 3) s=−1

(12.48)

Associated with each expansion coefficient is a pole factor; s, (s + 1), and (s + 3) are the pole factors associated with A1 , A2 ,

(12.50)

We note that (1) the presence of (s + 1) in the numerators of terms 1 and 3 on the left-hand side will force those terms to go to zero when evaluated at s = −1, (2) the middle term has only A2 in it, and (3) the reduction on the right-hand side of Eq. (12.50) eliminated the pole factor (s + 1) from the expression. Consequently, A2 =

(12.46)

The poles of F(s) are the values of s at which its denominator is zero. In the present case, the poles of F(s) are s = 0, s = −1, and s = −3. All three poles are real and distinct. By distinct we mean that no two or more poles are the same; in (s + 4)2 , for example, the pole s = −4 occurs twice, and therefore it is not distinct. F(s) can be decomposed into partial fractions corresponding to the three factors in the denominator of F(s): F(s) =

and A3 , respectively. To determine the value of any expansion coefficient we multiply both sides of Eq. (12.48) by the pole factor of that expansion coefficient, and then we evaluate them at s = pole value of that pole factor. The procedure is called the residue method. To determine A2 , for example, we multiply both sides of Eq. (12.48) by (s + 1), we reduce the expressions, and then we evaluate them at s = −1:     A1 A2 A3  (s + 1) + + s (s + 1) (s + 3) s=−1   (s + 1)(s2 − 4s + 3)  = . (12.49)  s(s + 1)(s + 3) s=−1

(−1)2 + 4 + 3 = −4. (−1)(−1 + 3)

Similarly, A1 = s F(s)|s=0

 s2 − 4s + 3  = = 1, (s + 1)(s + 3) s=0

and A3 = (s + 3) F(s)|s=−3 =

 s2 − 4s + 3  = 4. s(s + 1) s=−3

Having established the values of A1 to A3 , we now are ready to apply the inverse Laplace transform to Eq. (12.47): F(s) =

1 4 4 − + , s s+1 s+3

and with the help of Table 12-2, we obtain the result   1 4 4 f (t) = L −1 [F(s)] = L −1 − + s s+1 s+3 = [1 − 4e−t + 4e−3t ] u(t).

(12.51)

“book” — 2015/5/4 — 7:26 — page 646 — #17

646

CHAPTER 12

Building on this example, we can generalize the process to:

Distinct Real Poles

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

partial-fraction-expansion method is outlined by the following steps. Step 1. We are given a function F(s) composed of the product

Give a proper rational function defined by

F(s) = F1 (s) F2 (s),

N(s) N(s) F(s) = = , D(s) (s + p1 )(s + p2 ) . . . (s + pn ) (12.52) with numerator N(s) and distinct real poles −p1 to −pn , such that pi  = pj for all i  = j , F(s) can be expanded into the equivalent form:

=

n  i=1

Ai , s + pi

 A1 = (s + pi ) F(s)s=−p , i

i = 1, 2, . . . , n.

[F(s)]

= [A1 e−p1 t + A2 e−p2 t + · · · + An e−pn t ] u(t). (12.55)

Exercise 12-6: Apply the partial-fraction-expansion method to determine f (t), given that its Laplace transform is

F(s) =

F2 (s) =

12-5.2

(See

1 . (s + p)m

Bm B2 B1 + ··· + + . s + p (s + p)2 (s + p)m

(12.58)

)

Repeated Real Poles

We now consider the case when F(s) contains repeated real poles or a combination of distinct and repeated real poles. The

(12.59)

Step 3. Partial fraction expansion for the combination of the product F1 (s) F2 (s) is then given by F(s) =

10s + 16 . s(s + 2)(s + 4)

+ e−2t − 3e−4t ] u(t).

(12.57)

Step 2. Partial fraction representation for an m-repeated pole at s = −p consists of m terms:

= Answer: f (t) = [2

N(s) , (s + p1 )(s + p2 ) . . . (s + pn )

We note that F1 (s) is identical in form to Eq. (12.52) and contains only distinct real poles, −p1 to −pn , thereby qualifying it for representation by a series of terms as in Eq. (12.53). The second function, F2 (s), has an m-repeated pole at s = −p, where m is a positive integer. Also, the repeated pole is not a pole of F1 (s); p  = pi for i = 1, 2, . . . , n.

(12.54) In view of entry #3 in Table 12-2, the inverse Laplace transform or Eq. (12.53) is f (t) = L

F1 (s) =

(12.53)

with expansion coefficients A1 to An given by

−1

with

and

A1 A2 An + + ··· + s + p1 s + p2 s + pn

F(s) =

(12.56)

A2 An A1 + + ··· + s + p1 s + p2 s + pn B1 B2 Bm + + + ··· + 2 s + p (s + p) (s + p)m n  i=1

m

 Bj Ai + . s + pi (s + p)j

(12.60)

j =1

Step 4. Expansion coefficients A1 to An are determined by applying Eq. (12.54): Ai = (s + pi ) F(s)|s=−pi ,

i = 1, 2, . . . , n.

(12.61)

“book” — 2015/5/4 — 7:26 — page 647 — #18

12-5

PARTIAL FRACTION EXPANSION

647

Repeated Real Poles Expansion coefficients B1 to Bm are determined through a procedure that involves multiplication by (s + p)m , differentiation with respect to s, and evaluation at s = −p:  Bj =

  d m−j 1 m [(s + p) F(s)]  , m−j (m − j )! ds s=−p j = 1, 2, . . . , m.

(12.62)

Solution: In theory, any polynomial with real coefficients can be expressed as a product of linear and quadratic factors (of the form (s + p) and (s 2 + as + b), respectively). The process involves long division, but it requires knowledge of the roots of the polynomial, which can be determined through the application of numerical techniques. In the present case, a random check reveals that s = −2 and s = −3 are roots of D(s). Given that D(s) is fourth order, it should have four roots, including possible duplicates. Since s = −2 is a root of D(s), we should be able to factor out (s + 2) from it. Long division gives D(s) = s4 + 11s3 + 45s2 + 81s + 54

For the m, m − 1, and m − 2 terms, Eq. (12.62) reduces to: Bm = (s + p)m F(s)|s=−p ,    d m [(s + p) F(s)]  Bm−1 = , ds s=−p    1 d2 m  [(s + p) . F(s)] Bm−2 =  2 2! ds s=−p

(12.63a) (12.63b) (12.63c)

Thus, the evaluation of Bm does not involve any differentiation, that of Bm−1 involves differentiation with respect to s only once (and division by 1!), and that of Bm−2 involves differentiation twice and division by 2!. In practice, it is easiest to start by evaluating Bm first and then evaluating the other expansion coefficients in descending order. 5. Once the values of all of the expansion coefficients of Eq. (12.60) have been determined, transformation to the time domain is accomplished by applying entry #8 of Table 12-2, L −1



 (n − 1)! = t n−1 e−at u(t). (s + a)n

Next, we factor out (s + 3), which yields D(s) = (s + 2)(s + 3)(s2 + 6s + 9) = (s + 2)(s + 3)3 . Hence, F(s) has a distinct real pole at s = −2 and a triply repeated pole at s = −3, and the given expression can be rewritten as s2 + 3s + 3 . F(s) = (s + 2)(s + 3)3 Through partial fraction expansion, F(s) can be decomposed into F(s) =

B1 B3 B2 A + + + , 2 s + 2 s + 3 (s + 3) (s + 3)3

with A = (s + 2) F(s)|s=−2

(12.64)

f (t) = L−1 [F(s)] ⎡ ⎤ n m j −1   B t j =⎣ Ai e−pi t + e−pt ⎦ u(t). (12.65) (j − 1)!

= 1, s=−2

 s2 + 3s + 3  B3 = (s + 3) F(s)|s=−3 = = −3, s + 2 s=−3   d B2 = [(s + 3)3 F(s)] = 0, ds s=−3   1 d2 3  B1 = [(s + 3) F(s)] = −1.  2 ds2 s=−3

j =1

Hence, F(s) =

Example 12-3: Repeated Poles

Determine the inverse Laplace transform of F(s) =

 s2 + 3s + 3  = (s + 3)3 

3

The result is

i=1

= (s + 2)(s3 + 9s2 + 27s + 27).

N(s) s2 + 3s + 3 = 4 . 3 D(s) s + 11s + 45s2 + 81s + 54

1 1 3 − − , s + 2 s + 3 (s + 3)3

and use of Table 12-2 for the first two terms and application of Eq. (12.64) to the last term leads to   3 L−1 [F(s)] = e−2t − e−3t − t 2 e−3t u(t). 2

“book” — 2015/5/4 — 7:26 — page 648 — #19

648

TECHNOLOGY BRIEF 29: MAPPING THE ENTIRE WORLD IN 3-D

Technology Brief 29 Mapping the Entire World in 3-D Mapping software has become increasingly indispensable in the 21st-century industrialized world. Giving verbal directions to someone’s house has been replaced with directing them to Google Maps with the relevant address. Even more exciting and controversial, however, is the growing suite of 3-D virtual globe mapping software. Of these, arguably the most famous is the currently free Google Earth software. These packages allow the user to fly in virtual space around the world, into cities and remote areas and (in densely mapped areas) to view their own backyards, streets signs, and local landscapes; Google Sky includes virtual, navigable representations of the Moon, Mars, and the night sky.The tools are becoming an enabler for a new generation of armchair historians, archaeologists, demographers. They have already been used in search and rescue operations. How do these packages work? Where does this data come from? How is the world mapped?

Planes, Satellites, and Automobiles Data for these packages is acquired by specialized companies (including Google itself) that make use of satellites, aircraft, and (more recently) large fleets of specially equipped vans. The majority of data comes from several satellites orbiting earth financed by either national governments or private companies. For example, the U.S. National Aeronautics and Space Administration (NASA) has the long-standing Landsat 7 program which has 30-m imaging resolution and scans the earth in about 16 days. The European Space Agency’s ERS and Envisat satellites perform similar functions. All of these satellites perform functions other than visual spectrum imaging; some have infrared sensors, radar sensors, temperature sensors, etc. Several commercial satellites are now in orbit whose primary function is to map the globe in highresolution mode; DigitalGlobe’s WorldView-2 satellite, for example, provides 0.46 m spatial resolution—although not all data are publicly available. Most of these satellites maintain sun-synchronous orbits, which means that their orbits loop over or near the north and south poles and cross the equator twice on each loop. In this type of orbit, the satellite “visits” a given place at the same local time each visit, which is great for maintaining constant lighting

Figure TF29-1: The Shuttle Radar Topography Mission used an antenna located in the payload bay of the shuttle, and a second outboard antenna attached to the end of a 60-m mast. (Courtesy of NASA.) for satellite images. Additionally, 2-D visual information is supplemented with digital elevation model (DEM) data collected by NASA’s Shuttle Radar Topography Mission (SRTM). The SRTM (Fig. TF29-1) consisted of two radar antennas deployed on the space shuttle Endeavour during the 11-day mission of STS-99 in February 2000. A sample product is shown in Fig. TF29-2. Aircraft imaging complements the satellite data, although it is more expensive and available in limited areas. Several companies have launched fleets of specially equipped vans with multiple cameras, laser distance sensors, and on-board computation to collect, merge, and store high-resolution, three-dimensional data at street level. Additionally, Google Earth allows for userinputted 3-D information and models. Figure TF29-3 shows one such vehicle developed by TeleAtlas. Hundreds of similar vehicles roam the earth; the cameras provide images over 360◦ around the vehicle, and a laser system measures important distances like bridge and building heights; GPS tracking hardware records the vehicle’s position; and onboard computers synthesize everything and store it. As these vehicles visit more and more places, the 3-D map of the world continues to grow.

Imaging Software To produce a navigable, virtual representation of our globe, all of this data is then compiled, corrected, and

“book” — 2015/5/4 — 7:26 — page 649 — #20

TECHNOLOGY BRIEF 29: MAPPING THE ENTIRE WORLD IN 3-D

649

Figure TF29-2: A shaded relief image of Mount St. Helens in the state of Washington. (Courtesy of NASA.)

merged. This is not just a massive storage operation. Often, imagery comes from multiple sources that do not match exactly, there may be gaps between images and, very commonly, the color of the images must be corrected and made consistent. Fine-scale errors often are detectable with these map programs when data is incorrectly merged or have different dates; for example, pictures of a city might incorrectly show data from adjacent areas taken before and after major events, stitched together. Problems with incorporating 3D topographical data with the visual information are still common. For public-accessible programs, not all data is taken at the same time nor with the same frequency; for example, Google Earth guarantees that image data is no more than three years old. More expensive commercial software is often more timely. Beyond the compilation and merging of datasets, programs like Google Earth are increasingly integrating their software with both other software and mobile hardware. For example, Google Earth interfaces with both Wikipedia and the Google search engine as well as an increasing suite of information-providing programs. In a similar manner, some versions of commercial virtual

FigureTF29-3: A TeleAtlas van showing the imaging and laser equipment and the computation hardware inside the van.

globe programs can interface with GPS position-finding devices. Such programs take waypoints and tracks from the mobile GPS devices and merge them with available topographic, imaging, and other virtual globe datasets.

“book” — 2015/5/4 — 7:26 — page 650 — #21

650

CHAPTER 12

Concept Question 12-7: What purpose does the partial-

fraction-expansion method serve? (See

)

Concept Question 12-8: When evaluating the expansion coefficients of a function containing repeated poles, is it more practical to start by evaluating the coefficient of the fraction with the lowest-order pole or that with the highest-order pole? Why? (See )

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Determination of A, B1 , and B2 follows the same factormultiplication technique employed in Section 12-5.1 with  4s + 1  = −0.3, A = (s + 1) F(s)|s=−1 = 2 s + 4s + 13 s=−1 (12.70a) B1 = (s + 2 − j 3) F(s)|s=−2+j 3   4s + 1  = (s + 1)(s + 2 + j 3) s=−2+j 3 4(−2 + j 3) + 1 (−2 + j 3 + 1)(−2 + j 3 + 2 + j 3) −7 + j 12 ◦ = = 0.73e−j 78.2 , −18 − j 6 =

Exercise 12-7: Determine the inverse Laplace transform

of F(s) = Answer: f (t)

12-5.3

4s2

− 15s − 10 . (s + 2)3

= (18t 2 − 31t + 4)e−2t u(t).

and

(See

)

 In fact, the expansion coefficients associated with conjugate poles are always conjugate pairs themselves. 

The Laplace transform of a certain circuit is given by 4s + 1 . (s + 1)(s2 + 4s + 13)

(12.66)

In addition to the simple-pole factor, the denominator includes a quadratic-pole factor with roots s1 and s2 . Solution of s2 + 4s + 13 = 0 gives s1 = −2 + j 3,

s2 = −2 − j 3.

(12.67)

The fact that the two roots are complex conjugates of one another is a consequence of the property that for any physically realizable circuit, if it has any complex poles, those poles always appear in conjugate pairs. In view of Eq. (12.67), the quadratic factor is given by s2 + 4s + 13 = (s + 2 − j 3)(s + 2 + j 3),

B2 = (s + 2 + j 3) F(s)|s=−2−j 3   4s + 1 ◦  = = 0.73ej 78.2 . (12.70c)  (s + 1)(s + 2 − j 3) s=−2−j 3 We observe that B2 = B∗1 .

Distinct Complex Poles

F(s) =

(12.68)

and F(s) can now be expanded into partial fractions:

The inverse Laplace transform of Eq. (12.69) is  

−j 78.2◦ −1 −1 −0.3 −1 0.73e +L f (t) = L [F(s)] = L s+1 s + 2 − j3   ◦ 0.73ej 78.2 + L −1 s + 2 + j3 ◦

= [−0.3e−t + 0.73e−j 78.2 e−(2−j 3)t ◦

+ 0.73ej 78.2 e−(2+j 3)t ] u(t). (12.71) Because complex numbers do not belong in the time domain, our initial reaction to their presence in the solution given by Eq. (12.71) is that perhaps an error was committed somewhere along the way. The truth is the solution is correct, but incomplete. Terms 2 and 3 are conjugate pairs, so by applying Euler’s formula, they can be combined into a single term containing real quantities only: ◦

F(s) =

(12.70b)

A B1 B2 + + . s + 1 s + 2 − j3 s + 2 + j3



0.73e−j 78.2 e−(2−j 3)t + 0.73ej 78.2 e−(2+j 3)t (12.69)

Expansion coefficients B1 and B2 are printed in bold letters to signify the fact that they may be complex quantities.





= 0.73e−2t [ej (3t−78.2 ) + e−j (3t−78.2 ) ] = 2 × 0.73e−2t cos(3t − 78.2◦ ) = 1.46e−2t cos(3t − 78.2◦ ).

(12.72)

“book” — 2015/5/4 — 7:26 — page 651 — #22

12-5

PARTIAL FRACTION EXPANSION

Hence, the final time-domain solution is f (t) = [−0.3e−t + 1.46e−2t cos(3t − 78.2◦ )] u(t). (12.73)

Exercise 12-8: Determine the inverse Laplace transform

of

2s + 14 F(s) = 2 . s + 6s + 25 √

651 where B∗1 and B∗2 are the complex conjugates of B1 and B2 , respectively. Coefficients A, B1 , and B2 are evaluated as follows:  108(s2 + 2)  = 2, A = (s + 2) F(s)|s=−2 = 2 (s + 10s + 34)2  s=−2

B2 = (s + 5 + j 3) F(s)|s=−5−j 3   108(s2 + 2)  = 2 (s + 2)(s + 5 − j 3) s=−5−j 3 2

Answer: f (t) = [2 2 e−3t cos(4t − 45◦ )] u(t).

(See

=

)

108[(−5 − j 3)2 + 2] (−5 − j 3 + 2)(−5 − j 3 + 5 − j 3)2 ◦

= 24 + j 6 = 24.74ej 14 ,

12-5.4

Repeated Complex Poles

If the Laplace transform F(s) contains repeated complex poles, we can expand it into partial fractions by using a combination of the tools introduced in Sections 12-5.2 and 12-5.3. The process is illustrated in Example 12-4. Example 12-4: Five-Pole Function

Determine the inverse Laplace transform of F(s) =

108(s2 + 2) . (s + 2)(s2 + 10s + 34)2

and

  d 2 B1 = [(s + 5 + j 3) F(s)] ds s=−5−j 3   2  108(s + 2) d  = 2 ds (s + 2)(s + 5 − j 3) s=−5−j 3  108(2s) 108(s2 + 2) = − (s + 2)(s + 5 − j 3)2 (s + 2)2 (s + 5 − j 3)2   2 × 108(s2 + 2)  − 3 (s + 2)(s + 5 − j 3) s=−5−j 3 ◦

= −(1 + j 9) = 9.06e−j 96.34 . The remaining constants are

Solution: The roots of s2 + 10s + 34 = 0



B∗1 = 9.06ej 96.34 , and ◦

B∗2 = 24.74e−j 14 ,

are s1 = −5 − j 3 and s2 = −5 + j 3. Hence, F(s) =

108(s2 + 2) , (s + 2)(s + 5 + j 3)2 (s + 5 − j 3)2

and its partial fraction expansion can be expressed as F(s) =

B1 B2 A + + s + 2 s + 5 + j 3 (s + 5 + j 3)2 B∗1 B∗2 + + , s + 5 − j 3 (s + 5 − j 3)2

and the inverse Laplace transform is f (t) = L−1 [F(s)]  ◦ ◦ 2 9.06e−j 96.34 9.06ej 96.34 −1 =L + + s+2 s + 5 + j3 s + 5 − j3 ◦  ◦ j 14 −j 24.74e 14 24.74e + + (s + 5 + j 3)2 (s + 5 − j 3)2  −2t = 2e ◦



+ 9.06(e−j 96.34 e−(5+j 3)t + ej 96.34 e−(5−j 3)t )  ◦ ◦ + 24.74t (ej 14 e−(5+j 3)t + e−j 14 e−(5−j 3)t ) u(t) = [2e−2t + 18.12e−5t cos(3t + 96.34◦ ) + 49.48te−5t cos(3t − 14◦ )] u(t).

“book” — 2015/5/4 — 7:26 — page 652 — #23

652

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Table 12-3: Transform pairs for four types of poles. Pole

F(s)

f (t)

1. Distinct real

A s+a

Ae−at u(t)

2. Repeated real

A (s + a)n

 3. Distinct complex  4. Repeated complex

A

Aej θ Ae−j θ + s + a + jb s + a − jb



Aej θ Ae−j θ + (s + a + j b)n (s + a − j b)n

Example 12-5: Interesting Transform!

2Ae−at cos(bt − θ ) u(t) 

2At n−1 −at cos(bt − θ ) u(t) e (n − 1)!

By invoking property #3a of Table 6-4, we obtain the inverse Laplace transform

Determine the time-domain equivalent of the Laplace transform se−3s . F(s) = 2 s +4

f (t) = L

where F1 (s) =

s2

s s B1 B2 = = + , +4 (s + j 2)(s − j 2) s + j2 s − j2

−1



 1 e−3s 1 e−3s [F(s)] = L + 2 s + j2 2 s − j2   1 −j 2(t−3) = (e + ej 2(t−3) ) u(t − 3) 2 −1

= [cos(2t − 6)] u(t − 3).

Solution: We start by separating out the exponential e−3s from the remaining polynomial fraction. We do so by defining F(s) = e−3s F1 (s),

t n−1 e−at u(t) (n − 1)!

We conclude this section with Table 12-3, which lists F(s) and its corresponding inverse transform f (t) for all combinations of real versus complex, and distinct versus repeated, poles.

12-6

s-Domain Circuit Element Models

with B1 = (s + j 2) F(s)|s=−j 2

 s  −j 2 1 = = = ,  s − j 2 s=−j 2 −j 4 2

and B2 = B∗1 =

1 . 2

Hence, F(s) = e−3s F1 (s) =

e−3s e−3s + . 2(s + j 2) 2(s − j 2)

 The s-domain technique can be used to analyze circuits excited by sources with any type of variation— including pulse, step, ramp, sinusoid, and exponential— and provides a complete solution that incorporates both the steady state and transient components of the overall response.  We can apply the technique by transforming the differential equation associated with the circuit, or, equivalently, by transforming the circuit itself, which entails representing R, L, and C by s-domain models.

“book” — 2015/5/4 — 7:26 — page 653 — #24

12-6 S-DOMAIN CIRCUIT ELEMENT MODELS

12-6.1

653

Resistor in the s-Domain

12-6.3

Capacitor in the s-Domain

Similarly,

Application of the Laplace transform to Ohm’s law, L[υ] = L[Ri],

(12.74)

V = RI,

(12.75)

i=C

dυ dt

I = sCV − C υ(0− ),

(12.81)

leads to

where, by definition, V = L[υ],

I = L[i].

(12.76)

Hence, for the resistor the correspondence between the time and s-domains is υ = Ri

12-6.2

V = RI.

(12.77)

where υ(0− ) is the initial voltage across the capacitor. The s-domain circuit models for the capacitor are available in Table 12-4.  The s-domain transformation of circuit elements incorporates initial conditions associated with any energy storage that may have existed in capacitors and inductors at t = 0− . 

12-6.4

Impedances ZR , ZL , and ZC are defined in the s-domain in terms of voltage to current ratios under zero initial conditions [i(0− ) = υ(0− ) = 0]:

Inductor in the s-Domain

For R, the form of the i–υ relationship remained invariant under the transformation to the s-domain. That is not the case for L and C. Application of the Laplace transform to the i–υ relationship of the inductor,   di , L [υ] = L L dt

Impedance

(12.78)

ZR = R,

ZL = sL,

1 . sC

(12.82)

Exercise 12-9: Convert the circuit in Fig. E12.9 into the

s-domain.

iL

gives V = L[sI − i(0− )],

ZC =

and

+ _

(12.79)

where i(0− ) is the current that was flowing through the inductor at t = 0− . The time differentiation property (#6 in Table 12-1) was used in obtaining Eq. (12.79). The correspondence between the two domains is expressed as

L

υs(t)

R2 R1

C

Figure E12.9 Answer:

di dt

V = sLI − L i(0− ).

(12.80)

In the s-domain, an inductor is represented by an impedance ZL = sL, in series with a dc voltage source given by L i(0− ) or—through source transformation—in parallel with a dc current source i(0− )/s, as shown in Table 12-4. Note that the current I flows from (−) to (+) through the dc voltage source (if i(0− ) is positive).

sL + Vs _

(See

LiL(0−) _ +

υ=L

R1

)

R2 1 sC

CυC(0−)

“book” — 2015/5/4 — 7:26 — page 654 — #25

654

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Table 12-4: Circuit models for R, L, and C in the s-domain. Time-Domain

s-Domain

Resistor

i R

I υ

R

υ = Ri

V

V = RI

Inductor

IL iL L

IL

sL υL

diL dt t 1 υL dt + iL (0− ) iL = L

_ L iL(0−) +

VL

OR

sL

VL

iL(0−) s

υL = L

VL = sLIL − L iL (0− )

IL =

VL iL (0− ) + sL s

0−

Capacitor

IC

IC 1 sC

C

υC

(0−)

υC

s dυC dt t 1 iC dt + υC (0− ) υC = C

+

iC

VC

OR

1 sC

VC

C υC(0−)

_

iC = C

0−

VC =

IC υC (0− ) + sC s

IC = sCVC − C υC (0− )

“book” — 2015/5/4 — 7:26 — page 655 — #26

12-7 S-DOMAIN CIRCUIT ANALYSIS

655

12-7 s-Domain Circuit Analysis The circuit laws and analysis tools we used earlier in the time domain are equally applicable in the s-domain. They include KVL and KCL; voltage and current division; source transformation; source superposition; and Th´evenin and Norton equivalent circuits. Execution of the s-domain analysis technique entails the following four steps:

6A is(t) = 0 t=0

iR

iL

iC

R

L

C

t=1s

(a) RLC circuit in time domain

Solution Procedure: s-Domain Technique

V

Step 1: Evaluate the circuit at t = 0− to determine voltages across capacitors and currents through inductors. Use this information in conjunction with Table 12-4 to transform the circuit from the time domain to the s-domain. Step 2: Apply KVL, KCL, and the other circuit tools to obtain an explicit expression for the voltage or current of interest. Step 3: If necessary, expand the expression into partial fractions. Step 4: Use the list of transform pairs given in Tables 12-2 and 12-3 and the list of properties in Table 12-1 (if needed) to transform the partial fraction to the time domain. This process is illustrated through the next five examples.

Is

Determine the capacitor current response iC (t) to the rectangular pulse shown in Fig. 12-5(a), given that R = 125 , L = 0.1 H, and C = 4 mF. Solution: Per the proposed solution recipe, our first step should be to determine iL (0− ) and υC (0− ). In the present case, prior to activating the current source, the circuit contained no energy. Hence, iL (0− ) = 0 and υC (0− ) = 0, in which case transformation of the circuit elements to the s-domain entails replacing L with sL and C with 1/(sC). The s-domain circuit is shown in Fig. 12-5(b). The input source is is (t) = 6u(t) − 6u(t − 1),

(12.83)

and, according to entries #2 and #2a in Table 12-2, the s-domain expression for the source should be 6 6 Is = − e−s . s s

(12.84)

IL

R

sL

IC 1 sC

(b) s-domain Amps is(t)

6 4

iC(t)

2 0 −2

1

2

3

t (s)

−4 −6

Example 12-6: Parallel RLC Circuit

IR

(c) Source and capacitor currents Figure 12-5: Circuit for Example 12-6.

Our intermediate goal is to determine IC , the s-domain current through the capacitor in Fig. 12-5(b). Application of KCL at node V in Fig. 12-5(b) gives

1 1 + + sC = Is . (12.85) V R sL Also, since υC (0) = 0, IC is related to V by IC = sCV. Solution for IC leads to ⎛ ⎜ IC = ⎝

⎞ s2

⎟ ⎠ Is , s 1 2 s + + RC LC

(12.86)

“book” — 2015/5/4 — 7:26 — page 656 — #27

656

CHAPTER 12

where, in anticipation of applying partial-fraction expansion later on, we configured the denominator such that the coefficient of the highest-order s-term is 1. Since Is is the sum of two similar terms, we shall apply the superposition principle as follows: ⎛ ⎞ 2 −s s ⎜ ⎟ 6 6e − = IC1 + IC2 , IC = ⎝ s 1 ⎠ s s 2 s + + RC LC (12.87) where ⎛ ⎞ IC1

s2 6s ⎜ ⎟6 =⎝ , ⎠ = 1 1 s s s s2 + + + s2 + RC LC RC LC (12.88a)

and

⎛ IC2

⎜ =⎝

⎞ 6s ⎟ −s ⎠ (−e ). s 1 2 s + + RC LC

(12.88b)

Inserting the specified element values, namely R = 125 , L = 0.1 H, and C = 4 mF, leads to s2

6s . + 2s + 2500

The roots of s2 + 2s + 2500 = 0 are √ s1 = −1 − j 2499 ≈ −1 − j 50, √ s2 = −1 + j 2499 ≈ −1 + j 50.

(12.89)

(12.90a) (12.90b)

Hence, IC1 =

6s . (s + 1 + j 50)(s + 1 − j 50)

(12.91)

B = (s + 1 + j 50) = 3e

−j 1.15◦

B∗

B + , s + 1 + j 50 s + 1 − j 50

with

  6s  (s + 1 + j 50)(s + 1 − j 50) s=−1−j 50

.

(12.94)

Solution for IC2 : The expression for IC2 given by Eq. (12.88b) is identical to that for IC1 except for a multiplication factor of (−e−s ). Per the time-shift property in Table 12-1, iC2 (t) can be obtained from the expression for iC1 (t) by multiplying iC1 (t) by (−1) and delaying t by 1 s. Hence, iC2 (t) = −6e−(t−1) cos[50(t −1)+1.15◦ ] u(t −1) A. (12.95) Total solution: iC (t) = iC1 (t) + iC2 (t) − e−(t−1) cos[50(t − 1) + 1.15◦ ] u(t − 1)} A. (12.96) Figure 12-5(c) displays the waveforms of the source current is (t) and the capacitor current iC (t). Example 12-7: Two-State Power Supply

In the circuit shown in Fig. 12-6(a), the voltage source can operate at 125 V or 250 V, and when it switches between the two states, it does so gradually. Determine iL (t) for t ≥ 0, in response to  125 V for t < 0 (12.97) υs (t) = −2t (250 − 125e ) u(t) V for t ≥ 0.

iL (0− ) =

125 125 = = 10 mA, R 12.5 × 103

υC (0− ) = υL (0− ) = 0.

(12.92)

Hence, ◦

IC1

iC1 (t) = 6e−t cos(50t + 1.15◦ ) u(t) A.

Solution: The circuit condition at t < 0 is depicted in Fig. 12-6(b) wherein the capacitor and inductor have been replaced with an open circuit and short circuit, respectively. KVL leads to

Partial fraction expansion takes the form IC1 =

Per entry #3 in Table 12-3, the time-domain equivalent of IC1 is

= 6{e−t cos(50t + 1.15◦ ) u(t)

Solution for IC1 :

IC1 =

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM



3e−j 1.15 3ej 1.15 = + . (s + 1 + j 50) (s + 1 − j 50)

(12.93)

With the help of Table 12-2, the s-domain expression for Vs at t ≥ 0 is 250 125 Vs = − . (12.98) s s+2

“book” — 2015/5/4 — 7:26 — page 657 — #28

12-7 S-DOMAIN CIRCUIT ANALYSIS

657

υs

R = 12.5 kΩ

250 V

+ _

125 V

iL υs

C = 0.8 mF

L=2H

t

0

(a) Circuit 12.5 kΩ

125 V

iL(0−) = 10 mA

+ _

+ − _υC(0 ) = 0

(b) Circuit at t < 0 iL

12.5 kΩ

Vs

+ _

sL = 2s 1 1250 = sC s

I1

I2

_ +

L iL(0−) = 20 mV

(c) s-domain for t ≥ 0 iL (mA) 20

10

0 0

1

2 (d) iL(t)

Figure 12-6: Example 12-7.

3

t (s)

“book” — 2015/5/4 — 7:26 — page 658 — #29

658

CHAPTER 12

The s-domain circuit is shown in Fig. 12-6(c), in which L and C are represented by their s-domain models in accordance with Table 12-4, namely

IL sL

L

+

2s

=

20 mV + _

iL

L iL(0−) _

and

iC

C

− IC 1/sC υC(0 )/s +_

1250/s

=

.

By inspection, the mesh current equations for the two loops in Fig. 12-6(c) are

1250 1250 I1 − I2 = Vs , (12.99a) 12.5 × 103 + s s

1250 1250 − (12.99b) I1 + + 2s I2 = 20 × 10−3 . s s After replacing Vs with the expression given by Eq. (12.98), simultaneous solution of the two linear equations leads to 10s3 + 21s2 + 6252s + 25000 mA. I2 = s(s + 2)(s2 + 0.1s + 625)

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

The waveform of iL (t) is displayed in Fig. 12-6(d). Example 12-8: ac Source with dc Bias

The current source shown in the circuit of Fig. 12-7(a) is given by the displayed waveform, which consists of a 1.5 A dc source prior to t = 0, and the combination of a cosinusoidal waveform and a 1 A dc bias after t = 0. Determine υout (t) for t ≥ 0, given that R1 = 1 , R2 = 0.5 , and L = 0.5 H. Solution: The current source is characterized by  is (t) =

Since is (t) was nonzero before t = 0, we need to examine initial conditions by analyzing the circuit shown in Fig. 12-7(b), from which we deduce that iL (0− ) =

Is =

The roots of the quadratic term in the denominator are

s2 ≈ −0.05 + j 25.

10s3 + 21s2 + 6252s + 25000 . (12.101) s(s + 2)(s + 0.05 + j 25)(s + 0.05 − j 25)

Application of the partial-fractions expansion recipes outlined earlier in Section 12-5 leads to  ◦ ◦ 20 10 0.38e−j 90 0.38ej 90 − + + mA, I2 = s s + 2 s + 0.05 + j 25 s + 0.05 − j 25 and transformation to the time domain gives iL (t) = i2 (t)   = 20 − 10e−2t + 0.76e−0.05t cos(25t + 90◦ ) u(t) mA   = 20 − 10e−2t − 0.76e−0.05t sin 25t u(t) mA.

1 0.5s 1.5s2 + 16 + 2 = s s + 16 s(s2 + 16)

Figure 12-7(c) depicts the circuit in the s-domain, where we applied source transformation to convert (Is , R1 ) into a voltage source Vs = Is R1 , in series with R1 . At node Vout in the circuit of Fig. 12-7(c),

Hence, Eq. (12.100) can be rewritten as I2 =

1.5R1 1.5 × 1 = = 1 A. R1 + R 2 1 + 0.5

The s-domain expression for is (t) for t ≥ 0 is

(12.100)

s1 ≈ −0.05 − j 25,

1.5 A for t ≤ 0 (1 + 0.5 cos 4t) A for t ≥ 0.

Vout − Vs − LiL (0− ) Vout = 0, + R1 + sL R2 which gives Vout =

R2 [1.5s2 + 16 + s(s2 + 16)L iL (0− )] s(s2 + 16)[(R1 + R2 ) + sL]

=

s3 + 3s2 + 16s + 32 2s(s + 3)(s2 + 16)

=

s3 + 3s2 + 16s + 32 . 2s(s + 3)(s + j 4)(s − j 4)

Partial fraction expansion gives Vout =

A1 A2 B B∗ + + + , s s + 3 s + j4 s − j4

“book” — 2015/5/4 — 7:26 — page 659 — #30

12-7 S-DOMAIN CIRCUIT ANALYSIS

659

iL(0−) = 1 A

L

+ is(t)

R1

L

υout(t)

R2

R1

1.5 A

_

is(t)

R2

(b) At t = 0−

1.5 A ω = 4 rad/s

1A

sL

0.5 A

LiL(0−) _

t (a) Circuit and source waveform

+

Vout

+

R1 R2

+ Vs = R1Is _

Vout

_ (c) s-domain

υout (V) 1.0 0.75 0.5 0.33 0.25

υC(0)

0.05 V 0.05 V

t (s)

0

(d) υC(t) Figure 12-7: Example 12-8.

“book” — 2015/5/4 — 7:26 — page 660 — #31

660

CHAPTER 12

with

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Simultaneous solution of the two equations leads to A1 = s Vout |s=0

 s3 + 3s2 + 16s + 32  1 = = ,  2 2(s + 3)(s + 16) s=0 3

A2 = (s + 3) Vout |s=−3

 s3 + 3s2 + 16s + 32  =  2s(s2 + 16)

I1 =

= s=−3

8 , 75

I2 =

= s=−j 4

1 j 53◦ e . 20

8(s + 6) . + 29s + 40

(12.104b)

IL = I1 − I2 =

Hence, ◦



8 ej 53 e−j 53 1 + + + . 3s 75(s + 3) 20(s + j 4) 20(s − j 4)

The corresponding time-domain voltage is   8 −3t 1 1 + e cos(4t − 53◦ ) u(t) (V). + υout (t) = 3 75 10

IL =

A1 A2 A3 + + . s s + 1.85 s + 5.4

(12.106)

The values of A1 through A3 are obtained from 60 = 1.5, 40 A2 = (s + 1.85)IL |s=−1.85  4s2 + 29s + 60  = = −0.76, 4s(s + 5.4) s=−1.85 A1 = sIL |s=0 =

(12.107a)

(12.107b)

and

Example 12-9: Circuit with a Switch

Determine iL (t) in the circuit shown in Fig. 12-8(a) for t ≥ 0. Solution: We start by examining the state of the circuit at t = 0− (before closing the switch). Upon replacing L with a short circuit and C with an open circuit, as portrayed by the configuration in Fig. 12-8(b), we establish that υC (0− ) = 12 V.

4s2 + 29s + 60 4s2 + 29s + 60 = , (12.105) s(4s2 + 29s + 40) 4s(s + 1.85)(s + 5.4)

which can be represented by the partial fraction expansion

After the decay of the negative exponential term, the output becomes the sum of a dc term (1/3 V) and an ac term with an amplitude of 0.1 V. The circuit output response is displayed in Fig. 12-7(d).

and

4s2

The associated inductor current IL is

 s3 + 3s2 + 16s + 32  = 2s(s + 3)(s − j 4) 

iL (0− ) = 1 A

(12.104a)

and

B = (s + j 4) Vout |s=−j 4

Vout =

12s2 + 77s + 60 s(4s2 + 29s + 40)

(12.102)

For t ≥ 0, the s-domain equivalent of the original circuit is shown in Fig. 12-8(c), where we have replaced R2 with a short circuit, converted the dc source into its s-domain equivalent and, in accordance with the circuit models given in Table 12-4, converted L and C into impedances—each with its own appropriate voltage source. By inspection, the two mesh current equations are given by 24 + 2, (12.103a) (4 + 12 + 2s)I1 − (12 + 2s)I2 = s

5 12 − (12 + 2s)I1 + 12 + 2s + I2 = −2 − . (12.103b) s s

A3 = (s + 5.4)IL |s=−5.4 = 0.26.

(12.107c)

Hence, IL =

1.5 0.76 0.26 − + , s s + 1.85 s + 5.4

(12.108)

and the corresponding time-domain current for t ≥ 0 is iL (t) = [1.5 − 0.76e−1.85t + 0.26e−5.4t ] u(t) A.

(12.109)

The time variation of iL (t) is displayed in Fig. 12-8(d). Example 12-10: Op-Amp Integrator

The op-amp integrator circuit shown in Fig. 12-9(a) was first introduced in Section 5-6.1. We now examine its behavior by applying the s-domain analysis technique to a step-function input given by υi (t) = 10u(t) mV. The capacitor had no charge prior to t = 0 and the op-amp’s power supply voltage is Vcc = 10 V.

“book” — 2015/5/4 — 7:26 — page 661 — #32

12-7 S-DOMAIN CIRCUIT ANALYSIS

661

R2 t=0 R1

R2 iL

8Ω R1 V0

V0

4Ω + _

iL(0−) = 1 A



L 24 V

2H

C

4Ω L

+ _

24 V

R3 12 Ω

0.2 F

R3 12 Ω

C υC(0−) = 12 V

(b) Circuit at t = 0− iL

(a) Time-domain

I1 24 s

5 s

2s

4 + _

I2

_ +

2

+ _

12

12 s

(c) s-domain

iL (A) 2.0

1.5

1.0

iL(0)

t (s)

0 1

2

(d) IL(t) Figure 12-8: Circuit for Example 12-9.

Solution: The s-domain circuit is shown in Fig. 12-9(b), from which we deduce that Vout = −

ZC ZR

Vi = −

Vi 10−2 =− sR1 C1 s



50 s

=−

0.5 . s2

Application of entry #4 of Table 12-2 leads to υout1 (t) = −0.5t u(t). We observe in Fig. 12-9(c) that υout (t) is a negative ramp function that saturates at Vcc = −10 V when t reaches 20 s.

“book” — 2015/5/4 — 7:26 — page 662 — #33

662

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

1 2.5 ZC = sC =

4 μF υi 5 kΩ

1

_

ZR = 5

υout

+

Vi =

Vcc = 10 V

103

10−2

_

Vn = 0

s

(a) Circuit

105 s Vout

+ Vcc = 10 V

(b) s-domain

υi(t) 10 mV 2

4

6

8 10 12 14 16 18 20 22

2

4

6

8 10 12 14 16 18 20 22

t (s)

υout(t) t (s)

−2 V −4 V −6 V −8 V −10 V

saturation @ −Vcc

−12 V (c) υi(t) and υout(t)

Figure 12-9: Circuit for Example 12-10.

12-8 Multisim Analysis of Circuits Driven by Nontrivial Inputs

wide range of conditions and for a variety of different input waveforms.

The utility of SPICE simulators becomes most evident when trying to simulate circuits driven by nontrivial inputs—in contrast with sinusoids or dc voltages. In this section, we will revisit some of the examples we examined earlier in this chapter to demonstrate how easy it is to obtain solutions with Multisim and to compare the solutions with the analytical results based on the Laplace transform method. As a learning tool, Multisim is also very useful, in that it allows the user to test his/her understanding of core concepts by simulating circuits over a

Example 12-11: RC Circuit Response

A series RC circuit, with R = 500 k and C = 1 μF, is excited by a voltage source that delivers a 1 V, 1 s rectangular pulse. Draw the circuit in Multisim and generate the output response using the Transient Analysis tool. Solution: By now, we should be very familiar with how to create a pulse source in Multisim. The circuit is shown in

“book” — 2015/5/4 — 7:26 — page 663 — #34

12-8

MULTISIM ANALYSIS OF CIRCUITS DRIVEN BY NONTRIVIAL INPUTS

663

Rise segment and the second called the Fall segment, and this order is independent of whether the change in level is actually a rise or a fall. In the present case we need to simulate in the first segment an instantaneous change in level from 15 V down to 0 V. We do so by setting the Initial Value to 15 V, the Pulsed Value to 0 V, the Rise Delay Time to 0 s, and the Rise Time to 1 ns (which is practically the same as instantaneous). To

R1 = 2 Ω

(a) Circuit in Multisim

+ _

υin(t)

L=2H R2 = 5 Ω

+ R3 = 3 Ω

C = 0.1 F

a

υout(t)

_ b

(a) Time domain

(b) Response Figure 12-10: (a) RC circuit excited by a 1 V, 1 s rectangular pulse at 0.5 s, and (b) the corresponding response at node 2.

(b) Circuit in Multisim

Fig. 12-10(a), and the output response across the capacitor is displayed in Fig. 12-10(b). Note that a delay time of 0.5 s was introduced in the parameter selections of the pulse source in order to generate a clearer plot.

Input voltage Example 12-12: Interrupted Voltage Source in Multisim

Draw the circuit shown in Fig. 12-11(a) in Multisim and then use Transient Analysis to generate a plot of the voltage across the 3- resistor in response to an input excitation given by 15 V prior to t = 0, and 15(1 − e−2t ) V afterwards. Solution: The circuit is reproduced in Fig. 12-11(b). To model the exponential input voltage, we use the EXPONENTIAL VOLTAGE source which can accommodate both rising and falling exponentials. Multisim divides the exponential voltage into two segments, with the first called the

Output voltage (across R3)

(c) Response Figure 12-11: Multisim rendition of the circuit response to a sudden (but temporary) change in supply voltage level.

“book” — 2015/5/4 — 7:26 — page 664 — #35

664

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

(a) Time-voltage pairs for circuit

(b) Response Figure 12-12: Multisim rendition of the circuit response to an arbitrary input signal produced by the PWL (Piecewise Linear) source. simulate the second segment during which the voltage increases from 0 to 15V with a time constant of 0.5 s, we set the Fall Delay at 0 and the Fall Time Constant at 0.5 s. Applying Transient Analysis results in the responses displayed in Fig. 12-11(c). Example 12-13: Piecewise Linear Voltage Source

The PIECEWISE LINEAR VOLTAGE source allows you to define time-voltage pairs such that the source will be at a given voltage at the corresponding time and the source will “connect the dots” in between using a linear progression. Hence, entering the time-voltage pairs of (0,1), (1,1), and (2,4) will create a source which starts at 1 V and stays steady until 1 s, at which

time it will increase with a slope of 3 V/s to reach a value of 4 V at 2 s. Replace the Exponential voltage source in the circuit in Fig. 12-11(a) with a piecewise linear (PWL) voltage source with the time-voltage pairs: (0,1), (1,1), (2,4), (3,3), (7,−1), (8,5), and (9,5). Plot both the input and output in Transient Analysis from 0 to 10 s. Solution: With the PIECEWISE LINEAR VOLTAGE source in place, double-click on it and then make sure the Value tab is selected. Click on Enter data points in table, and insert the time-voltage pairs shown in Fig. 12-12(a). Applying Transient Analysis results in the responses displayed in Fig. 12-12(b).

“book” — 2015/5/4 — 7:26 — page 665 — #36

PROBLEMS

665

Summary Concepts • The Laplace-transform analysis technique transforms the circuit to a new domain, solves for the quantity of interest in that domain, and then transforms the solution back to the time domain. The technique can be applied to circuits with any type of excitation. • The Laplace transform has many useful properties

that can facilitate the process of finding the Laplace transform of a time function. • Under zero initial conditions, circuit elements R, L, and C transform to R, sL, and 1/sC, respectively, in the s-domain.

Mathematical and Physical Models Unit Impulse Function δ(t − T ) = 0

Time/s-Domain Equivalents

for t  = T

υ = Ri

Resistor

∞

di dt dυ i=C dt υ=L

Inductor

δ(t − T ) dt = 1 −∞

Capacitor

Laplace Transform ∞ F(s) = L[f (t)] =

V = RI V = sLI − L i(0− ) I = sCV − C υ(0− )

f (t) e−st dt

0−

Table 12-1 Tables 12-2 and 12-3

Properties Transform Pair

Important Terms

Provide definitions or explain the meaning of the following terms:

complex frequency convergence condition critically damped response damped natural frequency damping coefficient delay time delta function distinct domain transformation expansion coefficient Fall improper rational function

impulse function initial condition initial value inverse Laplace transform final condition Laplace transform Laplace transformation Laplace transform pair natural response one-sided transform overdamped response partial fraction expansion

PROBLEMS

pole pole factor proper rational function real region of convergence residue method resonant frequency Rise sampling property second-order circuit singularity function steady state response ∞ (a) G1 =

Sections 12-1 and 12-2: Impulse Response and Complex Algebra 12.1

Evaluate each of the following integrals.

step function time scaling transient response two-sided transform underdamped response uniqueness property unit impulse function unit step function universal property

(3t 3 + 2t 2 + 1)[δ(t) + 4δ(t − 2)] dt −∞

4 (b) G2 = −2

4(e−2t + 1)[δ(t) − 2δ(t − 2)] dt

“book” — 2015/5/4 — 7:26 — page 666 — #37

666

CHAPTER 12

f1(t)

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

f2(t) 2

4

0

2 t (s)

0 1

1

2

3

4

t (s)

−2

2 (a)

(b)

f3(t)

f4(t)

4

4

2

2 t (s)

0 1

3

t (s)

0

4

2

6

(c)

7

(d) f6(t)

f5(t)

10

10 2 t (s)

0

t (s)

0 2

−10

4

6

8

−10 (e)

(f) Figure P12.3: Waveforms for Problem 12.3.

20 (c) G3 =

Sections 12-2 and 12-4: Laplace Transform 3(t cos 2πt − 1)[δ(t) + δ(t − 10)] dt

12.3 Express each of the waveforms in Fig. P12.3 in terms of step functions and then determine its Laplace transform. [Recall that the ramp function is related to the step function by r(t − T ) = (t − T ) u(t − T ).] Assume that all waveforms are zero for t < 0.

−20

12.2

Evaluate each of the following integrals:

(a) G1 = (b) G2 = (c) G3 =

∞ 3 2 −∞ (3t − 4t + 3)[δ(t) + 2δ(t − 2)] dt. 4 −3t + 1)[δ(t) − 2δ(t − 2)] dt. −4 2(e 16 −12 4[t sin(2πt) − 1][δ(t − 1) + δ(t − 6)]

dt.

12.4 Determine the Laplace transform of each of the following functions by applying the properties given in Tables 12-1 and 12-2.

“book” — 2015/5/4 — 7:26 — page 667 — #38

PROBLEMS (a) f1 (t) = 4te−2t u(t) (b) f2 (t) = 10 cos(12t + 60◦ ) u(t) *(c) f3 (t) = 12e−3(t−4) u(t − 4) (d) f4 (t) = 30(e−3t + e3t ) u(t) (e) f5 (t) = 16e−2t cos 4t u(t) (f) f6 (t) = 20te−2t sin 4t u(t) 12.5 Determine the Laplace transform of each of the following functions by applying the properties given in Tables 12-1 and 12-2. *(a) h1 (t) = 12te−3(t−4) u(t − 4) (b) h2 (t) = 27t 2 sin(6t − 60◦ ) u(t) *(c) h3 (t) = 10t 3 e−2t u(t) (d) h4 (t) = 5(t − 6) u(t − 3) (e) h5 (t) = 10e−3t u(t − 4) (f) h6 (t) = 4e−2(t−3) u(t − 4) 12.6 Determine the Laplace transform of the following functions. (a) f1 (t) = 25 cos(4πt + 30◦ ) δ(t) (b) f2 (t) = 25 cos(4πt + 30◦ ) δ(t − 0.2) sin 3t (c) f3 (t) = 10 u(t) t d2 (d) f4 (t) = 2 [e−4t u(t)] dt d (e) f5 (t) = [4te−2t cos(4π t + 30◦ ) u(t)] dt (f) f6 (t) = e−3t cos(4t + 30◦ ) u(t) (g) f7 (t) = t 2 [u(t) − u(t − 4)] (h) f8 (t) = 10 cos(6πt + 30◦ ) δ(t − 0.2)

667 6 (s + 2)(s + 4) 4 (b) F2 (s) = (s + 1)(s + 2)2 (a) F1 (s) =

(c) F3 (s) =

3s3 + 36s2 + 131s + 144 s(s + 4)(s2 + 6s + 9)

(d) F4 (s) =

2s2 + 4s − 10 (s + 6)(s + 2)2

12.9 Obtain the inverse Laplace transform of each of the following functions. (a) F1 (s) =

s2 + 17s + 15 s(s2 + 6s + 5)

2s2 + 10s + 16 (s + 2)(s2 + 6s + 10) 4 (c) F3 (s) = (s + 2)3

*(b) F2 (s) =

(d) F4 (s) =

2(s3 + 12s2 + 16) (s + 1)(s + 4)3

12.10 Obtain the inverse Laplace transform of each of the following functions. (s + 2)2 s(s + 1)3 1 (b) F2 (s) = 2 (s + 4s + 5)2 √ 2(s + 1) *(c) F3 (s) = 2 s + 6s + 13 −2(s2 + 20) (d) F4 (s) = s(s2 + 8s + 20) (a) F1 (s) =

12.7 Determine the Laplace transform of each of the following functions: (a) f1 (t) = 2t 2 e−3t u(t) (b) f2 (t) = 5 sin(6t + 30◦ ) u(t) *(c) f3 (t) = 10e−4(t−3) u(t − 3) (d) f4 (t) = 15(e−2t − e2t ) u(t) (e) f5 (t) = 8e−4t cos(2t) u(t) (f) f6 (t) = 10te−t sin(2t) u(t)

12.11 Obtain the inverse Laplace transform of each of the following functions. 4(s − 4) (a) F1 (s) = 2 + 2 s + 16 4s 4 *(b) F2 (s) = + 2 s s +9 (s + 5)e−2s (c) F3 (s) = (s + 1)(s + 3)

Section 12-5: Partial Fractions

(d) F4 (s) =

(1 − e−4s )(24s + 40) (s + 2)(s + 10)

12.8 Obtain the inverse Laplace transform of each of the following functions by first applying the partial-fractionexpansion method. ∗

(e) F5 (s) =

s(s − 8)e−6s (s + 2)(s2 + 16)

(f) F6 (s) =

4s(2 − e−4s ) s2 + 9

Answer(s) available in Appendix G.

“book” — 2015/5/4 — 7:26 — page 668 — #39

668

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

Sections 12-6 and 12-7: s-Domain Analysis

i(t)

*12.12 In the circuit of Fig. P12.12(a), is (t) is given by the waveform shown in Fig. P12.12(b). Determine iL (t) for t ≥ 0, given that R1 = R2 = 2 and L = 4 H.

υs

R L

+ _

C

iL

R1 is(t)

(a) Circuit

L R2

υs 3V

(a) Circuit

2V 1V

is(t)

0

8A

1

t (s)

3

2

(b) υs(t)

8e−2t t

Figure P12.14: Circuit and waveform for Problem 12.14.

(b) is(t) Figure P12.12: Circuit and waveform for Problem 12.12.

R1 12.13 In the circuit of Fig. P12.13, υs (t) is given by υs (t) = [20u(t) − 5δ(t)] V. Determine υC (t) for t ≥ 0, given that L = 1 H, C = 0.5 F, and R = 6 .

υs(t)

iL

R2

+ _

L C

R υs(t)

+ _

L

C

+ _υC

(a) υs(t)

Figure P12.13: Circuit for Problem 12.13.

12.14 In the circuit of Fig. P12.14(a), υs (t) is given by the waveform displayed in Fig. P12.14(b). Determine i(t) for t ≥ 0, given that R = 2 , L = 6 H, and C = 2 F. *12.15 In the circuit of Fig. P12.15(a), υs (t) is given by the waveform displayed in Fig. P12.15(b). Determine iL (t), given that R1 = R2 = 4 , L = 2 H, and C = 0.5 F.

1V

1

2

3

4

t (s)

(b) Figure P12.15: Circuit and waveform for Problem 12.15.

“book” — 2015/5/4 — 7:26 — page 669 — #40

PROBLEMS 12.16

669 12.20 Determine υout (t) in the circuit in Fig. P12.20, given that υs (t) = 35u(t) V, υC1 (0− ) = 20 V, υC2 (0− ) = 0, R1 = 1 , C1 = 1 F, R2 = 0.5 , and C2 = 2 F.

In the circuit of Fig. P12.16, is (t) = 7u(t) + 4δ(t) (A).

Initially the capacitor had 8 J of energy stored in it. Determine υR1 (t) for t ≥ 0, given that R1 = 2 , R2 = 4 , and C = 1 F.

+ υs(t) _

+ υR1

_

is(t)

R1 C

12.17 In the circuit of Fig. P12.17, υs (t) is a rectangular 5 V pulse of duration 3 seconds starting at t = 0. Determine iL (t) for t ≥ 0, given that R1 = 6 , L = 1 H, and C = 13 F. Assume that initially the capacitor had no charge stored in it.

L

12.21 Determine iL (t) in the circuit of Fig. P12.21 for t ≥ 0, given that the switch was opened at t = 0 after it had been closed for a long time, υs = 12 mV, R0 = 5 , R1 = 10 , R2 = 20 , L = 0.2 H, and C = 6 mF.

R0

R1

R2 iL

t=0 L

C

Figure P12.21: Circuit for Problems 12.21 and 12.22.

t=0

C

*12.22 Repeat Problem 12.21, but assume that the switch had been open for a long time and then closed at t = 0. Set the dc source at 12 mV and the element values at R0 = 5 , R1 = 10 , R2 = 20 , L = 2 H, and C = 0.4 F.

Figure P12.17: Circuit for Problem 12.17.

*12.18 Determine υ(t) in the circuit of Fig. P12.18, given that υs (t) = 2u(t) V, R1 = 1 , R2 = 3 , C = 0.3689 F, and L = 0.2259 H.

12.23 Determine iL (t) in the circuit of Fig. P12.23, given that R1 = 1 , R2 = 6 , L = 1 H, and C = 0.5 F. Assume no energy was stored in the circuit segment to the right of the switch prior to t = 0.

R2

R1 + υs(t) _

R2 υout

_

+ 12 mV _

iL

+ _

+ C2

Figure P12.20: Circuit for Problem 12.20.

Figure P12.16: Circuit for Problem 12.16.

υs(t)

C1

+ _ υC

R2

R1

υC1

R1

iL(t) C

υ

L

R1 + 2V _

R2

iL

t=0 C

L

Figure P12.18: Circuit for Problems 12.18 and 12.19. Figure P12.23: Circuit for Problem 12.23.

12.19 Determine iL (t) in the circuit in Fig. P12.18, given that υs (t) = 2u(t), R1 = 2 , R2 = 6 , L = 2.215 H, and C = 0.0376 F.

12.24 Determine υC2 (t) in the circuit of Fig. P12.24, given that R = 200 , C1 = 1 mF, and C2 = 5 mF.

“book” — 2015/5/4 — 7:26 — page 670 — #41

670

CHAPTER 12

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

C1

R

R C

+ 50 δ(t) V _

C2

υC2

t=0

+ 7V _ L iL

Figure P12.24: Circuit for Problem 12.24.

Figure P12.27: Circuit for Problem 12.27.

*12.25 Determine iL (t) in the circuit of Fig. P12.25, given that before closing the switch υC (0− ) = 12 V. Also, the element values are R = 2 , L = 1.5 H, and C = 0.5 F.

*12.28 Apply mesh-current analysis in the s-domain to determine iL (t) in the circuit of Fig. P12.28, given that υs (t) = 44u(t) V, R1 = 2 , R2 = 4 , R3 = 6 , C = 0.1 F, and L = 4 H.

iL t=0

R

iL

L

L

R3

R1

υC + υs(t) _

R2

C

Figure P12.25: Circuit for Problem 12.25. Figure P12.28: Circuit for Problem 12.28.

12.26 Determine υout (t) in the circuit of Fig. P12.26, given that υs (t) = 11u(t) V, R1 = 2 , R2 = 4 , R3 = 6 , L = 1 H, and C = 0.5 F.

12.29 Determine υout (t) in the circuit of Fig. P12.29, given that υs (t) = 3u(t) V, R1 = 4 , R2 = 10 , and L = 2 H.

R3

+ υs(t) _

ix

C

R1

+ υs(t) _

+ L

2ix

R1

R2

L

+ R2

υout

_

υout

_

Figure P12.29: Circuit for Problems 12.29 and 12.30.

Figure P12.26: Circuit for Problem 12.26.

12.30 12.27 Determine iL (t) in the circuit of Fig. P12.27 for t ≥ 0, given that R = 4 , L = 1 H, and C = 0.5 F.

Repeat Problem 12.29 with υs (t) = 3δ(t) V.

*12.31 The voltage source in the circuit of Fig. P12.31 is, given by υs (t) = [10+5u(t)] V. Determine iL (t) for t ≥ 0, given that R1 = 1 , R2 = 1 , L = 2 H, and C = 1 F.

“book” — 2015/5/4 — 7:26 — page 671 — #42

PROBLEMS

671

R1

C

R2

iL

iL + υs(t) _

C

is(t)

L

R1

R2

L

Figure P12.36: Circuit for Problems 12.36 and 12.37. Figure P12.31: Circuit for Problems 12.31 and 12.35.

12.32 The current source in the circuit of Fig. P12.32 is given by is (t) = [10u(t) + 15δ(t)] mA. Determine υC (t) for t ≥ 0, given that R1 = 1 k , R2 = 1 k , and C = 2 mF.

*12.37 Given the current-source waveform displayed in Fig. P12.37, determine iL (t) in the circuit of Fig. P12.36, given that R1 = 10 , R2 = 5 , L = 0.6196 H, and LC = (1/15) s.

is(t)

R2 iC + is(t) _

R1

+ 6A

υC

C

6e−2t

_

0

t

0

Figure P12.32: Circuit for Problems 12.32 and 12.34.

Figure P12.37: Waveform for Problem 12.37.

12.33 The circuit in Fig. P12.33 is excited by a 10 V, 1 s long rectangular pulse. Determine i(t), given that R1 = 1 , R2 = 2 , and L = 1/3 H.

i(t)

R1 10 V

+ _

υs(t) = 0

12.38 If the circuit shown in Fig. P12.38(a) is excited by the current waveform is (t) shown in Fig. P12.38(b), determine i(t) for t ≥ 0, given that R1 = 10 , R2 = 5 , and C = 0.02 F.

i(t) L

R2

1s

is(t)

R1

Figure P12.33: Circuit for Problem 12.33.

*12.34 Repeat Problem 12.32 after replacing the current source with a 10 mA, 2 s long rectangular pulse. 12.35 Analyze the circuit shown in Fig. P12.31 to determine iL (t) in response to a voltage excitation υs (t) in the form of a 10 V rectangular pulse that starts at t = 0 and ends at t = 5 s. The element values are R1 = 1 , R2 = 3 , L = 2 H, and C = 0.5 F. 12.36 The current source in the circuit of Fig. P12.36 is given by is (t) = 6e−2t u(t) A. Determine iL (t) for t ≥ 0, given that R1 = 10 , R2 = 5 , L = 0.6196 H, and LC = (1/15) s.

C

R2

(a) Circuit is(t) 1.5 A ω = 4 rad/s

1A 0.5 A

t (b) Waveform Figure P12.38: Circuit for Problems 12.38 to 12.40.

“book” — 2015/5/4 — 7:26 — page 672 — #43

672

CHAPTER 12

12.39 If the circuit shown in Fig. P12.38(a) is excited by current waveform is (t) = 36te−6t u(t) mA, determine i(t) for t ≥ 0, given that R1 = 2 , R2 = 4 , and C = (1/8) F. 12.40 If the circuit shown in Fig. P12.38(a) is excited by a current waveform given by is (t) = 9te−3t u(t) mA, determine i(t) for t ≥ 0, given that R1 = 1 , R2 = 3 , and C = 1/3 F. 12.41 The circuit shown in Fig. P12.41 first was introduced in Problem 5.68. Then, a time-domain solution was sought for υout1 (t) and υout2 (t) for t ≥ 0, given that υi (t) = 10u(t) mV, VCC = 10 V for both op amps, and the two capacitors had no change prior to t = 0. Analyze the circuit and plot υout1 (t) and υout2 (t) using the Laplace transform technique.

4 μF υi 5 kΩ

_ +

_

υout2

+

Vcc = 10 V

*(a) υs (t) = 2u(t) (V), (b) υs (t) = 2 cos(1000t) u(t) (V), (c) υs (t) = 2e−t u(t) (V).

C1

R1 + _ υs(t)

Vcc = 10 V

12.42 Repeat Problem 12.41 retaining all element values and conditions but changing the input voltage to υi (t) = 0.4te−2t u(t). 12.43 For the circuit shown in Fig. P12.43, determine υout (t) given that R1 = 1 k , R2 = 4 k , and C = 1 μF, and (a) υs (t) = 2u(t) (V), (b) υs (t) = 2 cos(1000t) (V), (c) υs (t) = 2e−t u(t) (V).

R2

_

C

+ υs(t) _

+

+

+ υout(t)

C2

_

Figure P12.44: Op-amp circuit for Problem P12.44.

12.45 Apply a 1 V, 1 Hz signal with a 10 V dc offset to the circuit in Fig. P12.45 and plot both υC (t) and υR (t) for R = 1 and C = 1 F. Why are the dc offsets different for the voltages across the two components?

Figure P12.41: Circuit for Problems 12.41 and 12.42.

R1

_

R2

Section 12-8: Multisim Analysis of Circuits Driven by Nontrivial Inputs

5 μF υout1 1 MΩ

CIRCUIT ANALYSIS BY LAPLACE TRANSFORM

+ υout(t)

_

R

υs

+ _

+ υR _ C

+ υC _

Figure P12.45: Circuit for Problems 12.45 and 12.47.

12.46 Repeat Example 12-12, but vary the time constant of the exponent from 0.1 s to 5 s (pick more than 3 points). Plot all responses on the same display. 12.47 In Multisim, apply input signal υs (t) = 5t V to the circuit shown in Fig. P12.45 and plot both υC (t) and υR (t) for 0 to 5 s. Find the point at which υC (t) = υR (t). If we change the input signal to υs (t) = 10t V, will the point in time where υC (t) = υR (t) change? Explain. 12.48 Using a Piecewise Linear source in Multisim, build and simulate the circuit found in Fig. P12.33 (including the specified source). Plot i(t) for 0 to 2 s. Use R1 = 1 , R2 = 2 , and L = 1/3 H. Potpourri Questions

Figure P12.43: Op-amp circuit for Problem 12.43.

12.44 For the circuit shown in Fig. P12.44, determine υout (t) given that R1 = R2 = 100 , C1 = C2 = 1 μF, and

12.49 What techniques are used to generate the 3-D effect in 3-D TVs? 12.50 What type of imaging sensor is used to map the Earth surface in 3-D? How many antennas does it use?

“book” — 2015/5/4 — 7:26 — page 673 — #44

PROBLEMS

673

Integrative Problems: Analytical / Multisim / myDAQ To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available on .] m12.1 Step Response: For the circuit in Fig. m12.1: (a) Determine the transfer function H(s) = Vo (s)/Vs (s). Write the transfer function in simplified standard form with symbolic values. (b) Determine the output response υo (t) to the input υs (t) = 4u(t) by working in the Laplace domain. Assume the capacitor is initially discharged. (c) Plot υs (t) and υo (t) on the same graph from 0 to 5 ms using a tool such as MathScript or MATLAB for R = 5.6 k and C = 0.1 μF. Include a hard copy of the script used to create the plot. (d) Determine the following values for υo (t): (1) Initial value υo (0), (2) Time to reach 50% of the initial value, and (3) Final value.

R

R

υs

_ +

C R

Figure m12.1 Circuit for Problem m12.1.

υo

“book” — 2015/5/4 — 7:27 — page 674 — #1

13

CHAPTER

Fourier Analysis Technique τ=1s

Contents 13-1 13-2 TB30 13-3 13-4 TB31 13-5 TB32 13-6 13-7 13-8 13-9

Overview, 675 Fourier Series Analysis Technique, 675 Fourier Series Representation, 677 Bandwidth, Data Rate, and Communication, 688 Circuit Applications, 690 Average Power, 693 Synthetic Biology, 695 Fourier Transform, 697 Brain-Machine Interfaces, 702 Fourier Transform Pairs, 704 Fourier versus Laplace, 710 Circuit Analysis with Fourier Transform, 711 Multisim: Mixed-Signal Circuits and the Sigma-Delta Modulator, 713 Summary, 717 Problems, 718

Objectives Learn to: 

Express a periodic function in terms of a Fourier series using the cosine/sine, the amplitude/phase, and the complex exponential representations.



t (s) T = 10 s 2 |cn|

1.5

T = 10 s 1 0.5 0 −4

−2

0

2

4

nω0

A circuit driven by a periodic excitation can be analyzed by (a) representing the excitation in terms of a Fourier series composed of sinusoids, (b) applying phasor analysis to determine the output response due to each sinusoid, and (c) then adding all of the output responses to constitute the total response of the circuit.



Calculate the average power dissipated in or delivered by a component in a circuit excited by a periodic voltage or current.

Determine the line spectrum of a periodic waveform.



Evaluate the Fourier transform of a nonperiodic waveform.



Utilize symmetry consideratins in the evaluation of Fourier coefficients.



Apply the Fourier transform technique to analyze circuits excited by nonperiodic waveforms.



Explain the Gibbs phenomenon.





Analyze circuits excited by periodic waveforms.

Use Multisim to model the behavior of the SigmaDelta modulator.

“book” — 2015/5/4 — 7:27 — page 675 — #2

13-1

FOURIER SERIES ANALYSIS TECHNIQUE

675

Overview

L = (2/π) H

First introduced in Chapter 7, the phasor-domain analysis technique has proven to be a potent—and easy to implement— tool for determining the steady-state response of circuits excited by sinusoidal waveforms. As a periodic function with a period T , a sinusoidal signal shares a distinctive property with all other members of the family of periodic functions, namely the periodicity property given by Eq. (8.4) as

υs(t)

R=4Ω

υout(t)

_

(a) RL circuit υs (V) 3

Input x(t) = x(t + nT ),

+

+ _

T

(13.1)

where n is an integer. Given this natural connection between sinusoids and other periodic functions, can we somehow extend the phasor-domain solution technique to nonsinusoidal periodic excitations such as a square wave or a train of pulses? The answer is yes, and the process for realizing it is facilitated by two enabling mechanisms: the Fourier theorem and the superposition principle. The Fourier theorem makes it possible to mathematically characterize any periodic excitation in the form of a sum of multiple sinusoidal harmonics. The superposition principle allows us to apply phasor analysis to calculate the circuit response due to each harmonic as if it were the only excitation in the circuit and then to add all of the responses together, thereby realizing the response to the original periodic excitation. The first half of this chapter aims to demonstrate the mechanics of the solution process as well as to explain the physics associated with the circuit response to the different harmonics. The second half of the chapter is devoted to the Fourier transform, which is particularly useful for analyzing circuits excited by nonperiodic waveforms, such as single pulses or step functions. As we will see in Section 13-7, the Fourier transform is related to the Laplace transform of Chapter 12 and becomes identical to it under certain circumstances, but the two techniques are generally distinct (as are their conditions of applicability from the standpoint of circuit analysis).

1

−1

2

t (s)

−3 (b) Square-wave excitation υout (V) 4

Output

2 0 0

−1

1

2

3

t (s)

−2 −4 (c) Output response Figure 13-1: RL circuit excited by a square wave and corresponding output response.

Step 1: Express the periodic excitation in terms of Fourier harmonics

13-1 Fourier Series Analysis Technique By way of introducing the Fourier series analysis technique, let us consider the RL circuit shown in Fig. 13-1(a), which is excited by the square-wave voltage waveform shown in Fig. 13-1(b). The waveform amplitude is 3 V and its period T = 2 s. Our goal is to determine the output voltage response, υout (t). The solution procedure consists of three basic steps.

According to the Fourier theorem (which we will introduce and examine in detail in Section 11-2), the waveform shown in Fig. 13-1(b) can be represented by the series   12 1 1 υs (t) = cos ω0 t − cos 3ω0 t + cos 5ω0 t − · · · V, π 3 5 (13.2) where ω0 = 2π/T = 2π/2 = π (rad/s) is the fundamental angular frequency of the waveform. Since our present objective

“book” — 2015/5/4 — 7:27 — page 676 — #3

676

CHAPTER 13

is to outline the solution procedure, we will accept it as a given that the infinite-series representation given by Eq. (13.2) is indeed equivalent to the square wave of Fig. 13-1(b). The series consists of cosine functions of the form cos mω0 t with m assuming only odd values (1, 3, 5, etc.). Thus, the series contains only odd harmonics of ω0 . Other periodic waveforms may include both odd and even harmonics. The coefficient of the mth harmonic is equal to 1/m (relative to the coefficient of the fundamental), and its polarity is positive if m = 3, 7, . . . and negative if m = 5, 9, . . . . In view of these properties, we can replace m with (2n − 1) and cast υs (t) in the form ∞

1 12  υs (t) = cos[(2n − 1)π t] V. (13.3) (−1)n+1 π 2n − 1 n=1

In terms of its components, υs (t) is given by υs (t) = υs1 (t) + υs2 (t) + υs3 (t) + · · ·

(13.4)

with 12 cos ω0 t V, π 12 υs2 (t) = − cos 3ω0 t V, 3π υs1 (t) =

12 cos 5ω0 t V, 5π

etc.

For the circuit in Fig. 13-1(a), input voltage Vs1 acting alone would generate a corresponding output voltage Vout1 . Keeping in mind that Vs1 corresponds to υs1 (t) at ω = ω0 = π rad/s, voltage division gives    R  Vs Vout1 = R + j ωL ω=ω0 =π 1 =

4 4 + jπ ×

(for ω = ω0 ),

(13.5c)

(13.6)

(13.7b)

and Vs3 =

12 5π

V

(for ω = 5ω0 ),

etc.

(13.8)

υout1 (t) = Re[Vout1 ej ω0 t ] = 3.42 cos(ω0 t − 26.56◦ ) V. (13.9) Similarly, at ω = 3ω0 = 3π rad/s,   R  Vs Vout2 = R + j ωL ω=3ω0 =3π 2   4 12 = · − = −0.71 −56.31◦ V 3π 4 + j 3π × π2 (13.10)

υout2 (t) = Re[Vout2 ej 3ω0 t ] = −0.71 cos(3ω0 t − 56.31◦ ) V. (13.11) In view of the harmonic pattern expressed in the form of Eq. (13.3), for the harmonic at angular frequency ω = (2n − 1)ω0 = (2n − 1)π rad/s,

(13.7a)

(for ω = 3ω0 ),

12 = 3.42 −26.56◦ π

with a corresponding time-domain voltage

Voutn =

4 4 + j (2n − 1)π ×

= (−1)n+1

(13.7c)

Phasor voltages Vs1 , Vs2 , Vs3 , etc., are the counterparts of υs1 (t), υs2 (t), υs3 (t), etc., respectively.

2 π

· (−1)n+1

12 π(2n − 1)

24  π(2n − 1) 4 + (2n − 1)2

· ∠− tan 12 V π 12 V Vs2 = − 3π

·

(13.5b)

with Vs1 =

2 π

and

In the phasor domain, the counterpart of υs (t) is given by: Vs (t) = Vs1 (t) + Vs2 (t) + Vs3 (t) + · · ·

Step 2: Determine output responses to input harmonics

(13.5a)

and υs3 (t) =

FOURIER ANALYSIS TECHNIQUES

−1 [(2n−1)/2]

V.

(13.12)

The corresponding time domain voltage is υoutn (t) = Re[Voutn ej (2n−1)ω0 t ] 24  π(2n − 1) 4 + (2n − 1)2    2n − 1 · cos (2n − 1)ω0 t − tan−1 V, 2 (13.13)

= (−1)n+1

with ω0 = π rad/s.

“book” — 2015/5/4 — 7:27 — page 677 — #4

13-2

FOURIER SERIES REPRESENTATION

677

Step 3: Apply the superposition principle to determine υout (t) According to the superposition principle, if υout1 is the output generated by a linear circuit when excited by an input voltage υs1 acting alone and if similarly υout2 is the output due to υs2 acting alone, then the output due to the combination of υs1 and υs2 acting simultaneously is simply the sum of υout1 and υout2 . Moreover, the principle is extendable to any number of sources. In the present case, the square-wave excitation is equivalent to a series of sinusoidal sources υs1 , υs2 , . . . generating corresponding output voltages υout1 , υout2 , . . . . Consequently, υout (t) =

∞ 

υoutn (t)

n=1

=

∞ 

24  π(2n − 1) 4 + (2n − 1)2 n=1    −1 2n − 1 · cos (2n − 1)ω0 t − tan 2 (−1)n+1

Concept Question 13-3: What steps constitute the Fourier series solution procedure? (See )

13-2

Fourier Series Representation

In 1822, the French mathematician Jean-Baptiste Joseph Fourier developed an elegant formulation for representing periodic functions in terms of a series of sinusoidal harmonics. The representation is known today as the Fourier series, and the formulation is called the Fourier theorem. To guarantee that a periodic function f (t) has a realizable Fourier series, it should satisfy a set of conditions known as the Dirichlet conditions.  Fortunately, any periodic function generated by a real circuit will meet these conditions automatically and therefore, we are assured that its Fourier series does indeed exist.  The Fourier theorem states that a periodic function f (t) of period T can be cast in the form

= 3.42 cos(ω0 t − 26.56◦ ) − 0.71 cos(3ω0 t − 56.31◦ ) ◦

+ 0.28 cos(5ω0 t − 68.2 ) + · · · V,

f (t) = a0 + (13.14)

with ω0 = π rad/s. We note that the fundamental component of υout (t) has the dominant amplitude and that the higher the harmonic is, the smaller is its amplitude. This allows us to approximate υout (t) by retaining only a few terms, such as up to n = 10, depending on the level of desired accuracy. The plot of υout (t) displayed in Fig. 13-1(c) (which is based on only the first 10 terms) is sufficiently accurate for most practical applications. The foregoing three-step procedure (which is equally applicable to any linear circuit excited by any realistic periodic function) relied on the use of the Fourier theorem to express the square-wave pattern in terms of sinusoids. In the next section, we examine the attributes of the Fourier theorem and how we may apply it to any periodic function. Concept Question 13-1: The Fourier series technique

is applied to analyze circuits excited by what type of functions? (See ) Concept Question 13-2: How is the angular frequency

of the nth harmonic related to that of the fundamental ω0? How is ω0 related to the period T of the periodic function? (See )

∞ 

(an cos nω0 t + bn sin nω0 t),

n=1

(sine/cosine representation)

(13.15)

where ω0 , the fundamental angular frequency of f (t), is related to T by 2π ω0 = . (13.16) T The summation is an infinite series whose first pair of terms (for n = 1) involve cos ω0 t and sin ω0 t. Higher values of n involve sine and cosine functions at harmonic multiples of ω0 , namely 2ω0 , 3ω0 , etc. The constants a0 , an , and bn (for n = 1 to ∞) are collectively called the Fourier coefficients of f (t). Their values are determined by evaluating integral expressions involving f (t), namely, 1 a0 = T

2 an = T

T f (t) dt,

(13.17a)

f (t) cos nω0 t dt,

(13.17b)

0

T 0

“book” — 2015/5/4 — 7:27 — page 678 — #5

678

CHAPTER 13

FOURIER ANALYSIS TECHNIQUES

Table 13-1: Trigonometric integral properties for any

and 2 bn = T

integers m and n. The integration period T = 2π/ω0 .

T f (t) sin nω0 t dt.

(13.17c)

0

Property

Integral T sin nω0 t dt = 0

1

We will derive these expressions shortly.

0

 Even though the indicated limits of integration are from 0 to T , the expressions are equally valid if the lower limit is changed to t0 and the upper limit to (t0 +T ) for any value of t0 . In some cases, the evaluation is easier to perform by integrating from −T /2 to T /2. 

T cos nω0 t dt = 0

2 0

T sin nω0 t sin mω0 t dt = 0,

3

n = m

0

Coefficient a0 is equal to the time-average value of f (t). It is called the dc component of f (t), because the average values of the ac components are all zero.

T cos nω0 t cos mω0 t dt = 0,

4

n = m

0

T

13-2.1

Sine/Cosine Representation

sin nω0 t cos mω0 t dt = 0

5

To verify the validity of the expressions given by Eq. (13.17), we make use of the trigonometric integral properties listed in Table 13-1.

0

T sin2 nω0 t dt = T /2

6 0

dc Fourier component a0

T

Equation (8.5) in Chapter 8 states that the average value of a periodic function is obtained by integrating it over a complete period T and then dividing the integral by T . Applying the definition to Eq. (13.15) gives 1 T

T 0

1 f (t) dt = T 1 + T

1 = a0 + T 1 + T

T 0

T 0

T 0



∞ 

T a0 dt

cos2 nω0 t dt = T /2

7 0

Note: All integral properties remain valid when the arguments nω0 t and mω0 t are phase shifted by a constant angle φ0 . Thus, Property 1, for example, becomes 0T sin(nω0 t+φ0 ) dt = 0, and Property 5 becomes T 0 sin(nω0 t + φ0 ) cos(mω0 t + φ0 ) dt = 0.

0

an cos nω0 t + bn sin nω0 t

dt

(Property 2). Hence, all of the terms in Eq. (13.18) containing cos nω0 t or sin nω0 t will vanish, leaving behind

n=1

1 a1 cos ω0 t dt + T 1 b1 sin ω0 t dt + T

T a2 cos 2ω0 t dt + · · · 0

T b2 sin 2ω0 t dt + · · · .

1 T

T f (t) dt = a0 ,

(13.19)

0

which is identical to the definition given by Eq. (13.17a).

0

(13.18) According to Property 1 in Table 13-1, the average value of a sine function is zero, and the same is true for a cosine function

an Fourier coefficients Multiplication of both sides of Eq. (13.15) by cos mω0 t (with m being any integer value equal to or greater than 1), followed

“book” — 2015/5/4 — 7:27 — page 679 — #6

13-2

FOURIER SERIES REPRESENTATION

679

with integration over [0, T ] yields T

υ(t)

T f (t) cos mω0 t dt =

0

A

a0 cos mω0 t dt

0

0

+

T  ∞

−T/2 an cos nω0 t cos mω0 t dt

+

−A

(a) Original waveform

0 n=1

T  ∞

t T/2

υ(t) bn sin nω0 t cos mω0 t dt.

A

(13.20)

0 n=1

0 −T/2

On the right-hand side of Eq. (13.20): (1) The term containing a0 is equal to zero (Property 2 in Table 13-1).

(3) All terms containing an are equal to zero (Property 4), except when m = n, in which case Property 7 applies. Hence, after eliminating all of the zero-valued terms and then setting m = n in the two remaining terms, we have f (t) cos nω0 t dt = an

T , 2

−A υ(t) A 0

−T/2

T/2

−A

t

(c) Fourier series with 3 terms υ(t) A

(13.21)

0

t T/2

(b) First term of Fourier series

(2) All terms containing bn are equal to zero (Property 5).

T

Fourier series with only 1 term

0 −T/2

which proves Eq. (13.17b).

T/2

−A

t

(d) Fourier series with 10 terms bn Fourier coefficients Similarly, if we were to repeat the preceding process, after multiplication of Eq. (13.15) by sin mω0 t (instead of cos mω0 t), we would conclude with a result affirming the validity of Eq. (13.17c). To develop an appreciation for how the components of the Fourier series add up to represent the periodic waveform, let us consider the square-wave voltage waveform shown in Fig. 13-2(a). Over the period extending from −T /2 to T /2, υ(t) is given by ⎧ ⎪ ⎨−A, for − T /2 < t < −T /4, υ(t) = A, for − T /4 < t < T /4, ⎪ ⎩ −A, for T /4 < t < −T /2. If we apply Eq. (13.17)—with integration limits [−T /2, T /2]—to evaluate the Fourier coefficients and then use them in Eq. (13.15), we end up with the series

υ(t) A 0 −T/2

T/2

−A

t

(e) Fourier series with 100 terms Figure 13-2: Comparison of the square-wave waveform with its Fourier series representation using only the first term (b), the sum of the first three (c), ten (d), and 100 terms (e).

 2nπ t nπ 2 T n=1     4A 2π t 4A 6π t = cos − cos π T 3π T   4A 10π t + cos − ··· . 5π T

υ(t) =

∞  4A

sin

 nπ 



cos

“book” — 2015/5/4 — 7:27 — page 680 — #7

680

CHAPTER 13

Alone, the first term of the series provides a crude approximation of the square wave (Fig. 13-2(b)), but as we add more and more terms, the sum starts to better resemble the general shape of the square wave, as demonstrated by the waveforms in Figs. 13-2(c) to (e).

f(t) 20 (a) Original

Example 13-1: Sawtooth Waveform

−4

Express the sawtooth waveform shown in Fig. 13-3(a) in terms of a Fourier series, and then evaluate how well the original waveform is represented by a truncated series in which the summation stops when n reaches a specified truncation number nmax . Generate plots for nmax = 1, 2, 10, and 100. Solution: The sawtooth waveform is characterized by a period T = 4 s and ω0 = 2π/T = π/2 (rad/s). Over the waveform’s first cycle (t = 0 to t = 4 s), its amplitude variation is given by f (t) = 5t

2 an = T 2 = 4

T

0 f(t)

4

8

4

8

4

8

4

8

(b) nmax = 1

4

8

10

−4

(c) nmax = 2

0 f(t) 20

0

10

−4

4 5t dt = 10, 0

(d) nmax = 10

0 f(t) 20

f (t) cos(nω0 t) dt

−4

0

 nπ  t dt = 0, 2

0

t (s)

10

T

5t cos

t (s)

(for 0 ≤ t ≤ 4 s).

1 f (t) dt = 4

4

t (s)

20

Application of Eq. (13.17) yields: 1 a0 = T

FOURIER ANALYSIS TECHNIQUES

0 f(t)

t (s)

20 (e) nmax = 100

10

and 2 bn = T 2 = 4

T f (t) sin(nω0 t) dt

0

0

t (s)

Figure 13-3: Sawtooth waveform: (a) original waveform,

0

4

−4

 nπ  20 5t sin t dt = − . 2 nπ

Upon inserting these results into Eq. (13.15), we obtain the following complete Fourier series representation for the sawtooth waveform: ∞  nπ  20  1 sin t . f (t) = 10 − π n 2 n=1

(b)–(e) representation by a truncated Fourier series with nmax = 1, 2, 10, and 100, respectively.

 The nmax -truncated series is identical in form to the complete series, except that the summation is terminated after the index n reaches nmax .  Figures 13-3(b) through (e) display the waveforms calculated using the truncated series with nmax = 1, 2, 10, and 100. As expected, the addition of more terms improves the accuracy of

“book” — 2015/5/4 — 7:27 — page 681 — #8

13-2

FOURIER SERIES REPRESENTATION

681

the Fourier series representation, but even with only 10 terms (in addition to the dc component), the truncated series appears to provide a reasonable approximation of the original waveform.

Thus, an cos nω0 t + bn sin nω0 t = An cos φn cos nω0 t − An sin φn sin nω0 t. (13.24)

Concept Question 13-4: Is the Fourier series representa-

tion given by Eq. (13.15) applicable to a periodic function that starts at t = 0 (and is zero for t < 0)? (See ) Concept Question 13-5: What is a truncated series?

(See

)

Upon equating the coefficients of cos nω0 t and sin nω0 t on one side of the equation to their respective counterparts on the other side, we have an = An cos φn

Exercise 13-1: Obtain the Fourier series representation

An =

0 0

−2

(13.25)



an2 + bn2

and

⎧   ⎪ −1 bn ⎪ an > 0, − tan ⎪ ⎪ ⎨ an φn =   ⎪ bn ⎪ −1 ⎪ ⎪ an < 0. π − tan ⎩ an

10

−4

bn = −An sin φn ,

which can be combined to yield the relationships

for the waveform shown in Fig. E13.1.

f(t)

and

2

4

t (s)

(13.26)

−10 In complex form,

Figure E13.1

An φn = an − j bn .

Answer:

f (t) =

∞   20 nπ t (1 − cos nπ ) cos n2 π 2 2 n=1

+

(See

(13.27)

In view of Eq. (13.22), the cosine/sine Fourier series representation of f (t) can be rewritten in the alternative amplitude/phase format

 nπ t 10 (1 − cos nπ ) sin . nπ 2

f (t) = a0 +

)

∞ 

An cos(nω0 t + φn ).

(13.28)

n=1

13-2.2 Amplitude and Phase Representation In the sine/cosine Fourier series representation given by Eq. (13.15), at each value of the integer index n, the summation contains the sum of a sine term and a cosine term, with both at angular frequency nω0 . The sum can be converted into a single sinusoid as follows. For n ≥ 0, an cos nω0 t + bn sin nω0 t = An cos(nω0 t + φn ),

(13.22)

where An is called the amplitude of the nth harmonic and φn is its associated phase. The relationships between (An , φn ) and (an , bn ) are obtained by expanding the right-hand side of Eq. (13.22) in accordance with the trigonometric identity cos(x + y) = cos x cos y − sin x sin y.

(13.23)

(amplitude/phase representation) Associated with each discrete frequency harmonic nω0 is an amplitude An and a phase φn . A line spectrum of a periodic signal f (t) is a visual depiction of its Fourier coefficients, An and φn . Its amplitude spectrum consists of vertical lines located at discrete values along the ω axis, with a line of height a0 located at dc (ω = 0), another of height A1 at ω = ω0 , a third of height A2 at ω = 2ω0 , and so on. Similarly, the phase spectrum of f (t) consists of lines of lengths proportional to the values of φn with each located at its corresponding harmonic nω0 . Line spectra show at a glance which frequencies in the spectrum of f (t) are most significant and which are not. Example 13-2 provides an illustration.

“book” — 2015/5/4 — 7:27 — page 682 — #9

682

CHAPTER 13

f(t)

The dc component of f (t) is given by

1

−2

0 2 0 1 (a) Periodic waveform

−1

1 a0 = T

3

t (s)

An 0.38

0.4

an =

A1 0.16 A2 0.11 0.08 0.06 A3 A4 A5

0 ω0

2ω0

3ω0

4ω0

5ω0

2 T

2 = 2

1 f (t) dt = 2

0

1 (1 − t) dt = 0.25, 0

T f (t) cos nω0 t dt 0

1 (1 − t) cos nπ t dt 0

ω

(b) Amplitude spectrum

=

1  1   1 t 1  sin nπ t  − cos nπ t + sin nπ t  2 2 nπ n π nπ 0 0

=

1 [1 − cos nπ ] n2 π 2

and

ϕn 90o

2 bn = T

ω0

2ω0

ϕ1 −90o

T

which is equal to the area under a single triangle, divided by the period T = 2 s. For the other Fourier coefficients, evaluation of the expressions given by Eqs. (13.17b and c) leads to

a0

0.25 0.2

FOURIER ANALYSIS TECHNIQUES

ϕ2

3ω0 ϕ3

4ω0 ϕ4

5ω0

=−

ω

ϕ5

T

2 f (t) sin nω0 t dt = 2

0

1 (1 − t) sin nπ t dt 0

1  1   1 t 1  sin nπ t − cos nπ t  − cos nπ t  2 2 nπ n π nπ 0 0

1 . nπ By Eq. (13.26), the harmonic amplitudes and phases are given by

 2  2 1/2  1 1 + 2 An = an + bn2 = [1 − cos nπ] + n2 π 2 nπ ⎧ 1/2 1 4 ⎪ ⎪ ⎪ + for n = odd ⎨ n4 π 4 n2 π 2 = ⎪ ⎪ ⎪ ⎩ 1 for n = even nπ =

−57.5o

o

−90o −78 −90o (c) Phase spectrum

−83o

Figure 13-4: Periodic waveform of Example 13-2 with its associated line spectra.

Example 13-2: Line Spectra

Generate and plot the amplitude and phase spectra of the periodic waveform displayed in Fig. 13-4(a). Solution: The periodic waveform has a period T = 2 s. Hence, ω0 = 2π/T = 2π/2 = π rad/s, and the functional expression for f (t) over its first cycle along the positive t axis is  1 − t for 0 < t ≤ 1 s, f (t) = 0 for 1 ≤ t ≤ 2 s.

and φn = − tan−1

  bn nπ = − tan−1 an [1 − cos nπ]  nπ   − tan−1 for n = odd, 2 = ◦ for n = even. −90

“book” — 2015/5/4 — 7:27 — page 683 — #10

13-2

FOURIER SERIES REPRESENTATION

683

The values of An and φn for the first three terms are A1 = 0.38,

φ1 = −57.5◦ ,

A2 = 0.16,

φ2 = −90◦ ,

A3 = 0.11,

φ3 = −78◦ .

Spectral plots of An and φn are shown in Figs. 13-4(b) and (c), respectively. Exercise 13-2: Obtain the line spectra associated with the

periodic function of Exercise 13-1. Answer:

 20 An = [1 − cos(nπ )] 2 2 n π and φn = − tan−1

(See

13-2.3

1+

 nπ  2

The waveforms displayed in Figs. 13-5(b) and (c) exhibit even symmetry, as do the waveforms of sin2 ωt and | sin ωt|, among many others. In contrast, the sine and square waves shown in Fig. 13-5(d) and (e) exhibit odd symmetry; in each case, the shape of the waveform on the left-hand side of the vertical axis is the inverted mirror image of the waveform on the right-hand side. Thus, for an odd function; f (t) = −f (−t)

(13.30)

In the case of the square wave, were we to shift the waveform by T /4 to the left, it would switch from an odd function into an even function.

13-2.4

n2 π 2 4

(odd symmetry).

Even-Function Fourier Coefficients

Even symmetry allows us to simplify Eq. (13.17) to the following expressions:

.

Even Symmetry: f (t) = f (−t)

) 2 a0 = T

Symmetry Considerations

According to Eq. (13.17), determination of the Fourier coefficients involves evaluation of three definite integrals involving f (t). If f (t) exhibits symmetry properties, the evaluation process can be simplified significantly. dc Symmetry Since a0 is equal to the average value of f (t) which is proportional to the net area under the waveform over the span of a complete cycle, it follows that a0 = 0 if the waveform is such that the area above the zero horizontal axis is equal to the area below it. The waveforms in Figs. 13-5(a), (b), (d), and (e) are examples of dc-symmetrical functions. Even and odd symmetry The waveform of a function f (t) possesses even symmetry if it is symmetrical with respect to the vertical axis; the shape of the waveform on the left-hand side of the vertical axis is the mirror image of the waveform on the right-hand side. Mathematically, an even function satisfies the condition f (t) = f (−t)

(even symmetry).

(13.29)

4 an = T

T /2 f (t) dt, 0

T /2 f (t) cos(nω0 t) dt,

(13.31)

0

bn = 0, An = |an |,

 and

φn =

0 180◦

if an > 0, if an < 0.

The expressions for a0 and an are the same as given earlier by Eqs. (13.17a and b), except that the integration limits are now over half of a period and the integral has been multiplied by a factor of 2. The simplification is justified by the even symmetry of f (t). As was stated in connection with Eq. (13.17), the only restriction associated with the integration limits is that the upper limit has to be greater than the lower limit by exactly T . Hence, by choosing the limits to be [−T /2, T /2], and then recognizing that the integral of f (t) over [−T /2, 0] is equal to the integral over [0, T /2], we justify the changes reflected in the expression for a0 . A similar argument applies to the expression for an based on the fact that multiplication of an even function f (t) by cos nω0 t (which itself is an even function) yields an even function. The rationale for setting bn = 0 for all n

“book” — 2015/5/4 — 7:27 — page 684 — #11

684

CHAPTER 13

f(t)

f(t)

dc symmetry only

T/2

0 0

−T

2T T

0

t

0

−T/2

T/2

t

T

−A

−A (a) dc symmetry f(t)

(b) cos (2πt/T) f(t)

Even symmetry only

A

−T/2

dc and even symmetry

A

A

−T

FOURIER ANALYSIS TECHNIQUES

0 0

A

T/2

T

dc and odd symmetry

0

t

0

−T/2

(c) Even symmetry

T/2

t

T

−A (d) sin (2πt/T) f(t)

Line of even symmetry (equal on both sides) and odd symmetry (equal and opposite on both sides) t

dc and odd symmetry

A −T

Line of dc symmetry

−T/2

T/2

0

T

t

0 −A (e) odd symmetry

Figure 13-5: Waveforms with (a) dc symmetry, (b and c) even symmetry, and (d and e) odd symmetry.

relies on the fact that multiplication of an even function f (t) by sin nω0 t (which is an odd function) yields an odd function, and integration of an odd function over [−T /2, T /2] is always equal to zero. This is because the integral of an odd function over [−T /2, 0] is equal in magnitude but opposite in sign to the integral over [0, T /2].

13-2.5

Odd-Function Fourier Coefficients

In view of the preceding discussion, it follows that for a function with odd symmetry we have the following equations:

Odd Symmetry: f (t) = −f (−t) a0 = 0, 4 bn = T

an = 0, T /2 f (t) sin(nω0 t) dt, 0

An = |bn |

 and

φn =

−90◦ 90◦

(13.32) if bn > 0, if bn < 0.

Selected waveforms are displayed in Table 13-2, together with their corresponding Fourier series expressions.

“book” — 2015/5/4 — 7:27 — page 685 — #12

13-2

FOURIER SERIES REPRESENTATION

685

Table 13-2: Fourier series expressions for a select set of periodic waveforms. Waveform

A 0

1. Square Wave

T/2

−T −T/2 −A 2. Time-Shifted Square Wave

T/2

−T −T/2 A

4. Triangular Wave

−T/2

τ

−T/2

6. Sawtooth

−T

f (t) =

t

f (t) =

t

T

f (t) =

7. Backward Sawtooth

−2T −T A

8. Full-Wave Rectified Sinusoid

A −T/2

t

2T

T

2T

t

T/2

T

3T/2

Example 13-3: M-Periodic Waveform

Evaluate the Fourier coefficients of the M-periodic waveform shown in Fig. 13-6(a),

(−1)n+1

  2nπ t 2A sin nπ T

  ∞ A  A 2nπ t + sin 2 nπ T   ∞ 2A  4A 2nπ t + cos π T π(1 − 4n2 ) n=1

f (t) =

0

   nπ  2nπ t 8A sin sin 2 T n2 π 2

n=1

f (t) =

0 f(t)

−T 9. Half-Wave Rectified Sinusoid

0 T f(t)

∞ 

  2nπ t 8A cos T n2 π 2

n=1

f(t)

A

∞  n=1 n=odd

f(t)

−A

∞  n=1 n=odd

f(t)

−A

  ∞  nπ τ   Aτ 2A 2nπ t + sin cos T nπ T T n=1

f (t) =

t

T/2

A 0

f (t) =

f(t)

−A A 0

t

T

T/2

  ∞  4A 2nπ t sin nπ T

n=1 n=odd

−A f(t)

A 0

  ∞  nπ   4A 2nπ t sin cos nπ 2 T

n=1

f (t) =

t

T

0 0

−T

t

T

f (t) =

f(t)

A 0

3. Pulse Train

5. Shifted Triangular Wave

Fourier Series

f(t)

t

    ∞  A A 2π t 2A 2nπ t + sin + cos π 2 T T π(1 − n2 ) n=2 n=even

Solution: The M waveform is even-symmetrical, its period is T = 4 s, ω0 = 2π/T = π/2 rad/s, and its functional form over the positive half period is:  1 (1 + t) 0 ≤ t ≤ 1 s, f (t) = 2 0 1 ≤ t ≤ 2 s.

“book” — 2015/5/4 — 7:27 — page 686 — #13

686

CHAPTER 13

f(t)

4

5

t

0.5

0 −0.2

An 0.5 0.4 0.375 0.3

0.43 A1

a0

5 terms

1

1 0.5 0 −5 −4 −3 −2 −1 0 1 2 3 (a) M waveform

Amplitude

−5 −4 −3 −2 −1 0 1 (d) nmax = 5

0.23

0.2 0.1 A2

0.1

A3

0 ω0

2ω0 3ω0 4ω0 (b) Amplitude spectrum

ϕn 180o

180o

180o

ϕ2

ϕ3

5ω0

ω

2

3

4

5

2

3

4

5

2

3

4

5

10 terms

1

0.12 A5

A4 = 0

FOURIER ANALYSIS TECHNIQUES

0.5

0 −0.2

Phase

−5 −4 −3 −2 −1

0

1

(e) nmax = 10 ϕ1 = 0

0

ω0

ϕ4 = 0 ϕ5 = 0

2ω0 3ω0 4ω0 (c) Phase spectrum

5ω0

ω

1000 terms

1

0.5

0 −0.2

Gibbs phenomenon

−5 −4 −3 −2 −1

0

1

(f) nmax = 1000 Figure 13-6: Plots for Example 13-3. Application of Eq. (13.31) yields:

4 = 4

T /2 1 1 2 2 f (t) dt = (1 + t) dt = 0.375, a0 = T 4 2 0

an =

4 T

f (t) cos nω0 t dt 0

0

1 (1 + t) cos nω0 t dt 2

 2 nπ 2  nπ = sin + 2 2 cos −1 , nπ 2 n π 2

0

T /2

1

and bn = 0.

“book” — 2015/5/4 — 7:27 — page 687 — #14

13-2

FOURIER SERIES REPRESENTATION

687

Since bn = 0, we have for n  = 0

f1(t)

 An = |an |

φn =

and

0 180◦

A

if an > 0, if an < 0.

0 −T/2

T/2

−A

Figures 13-6(b) and (c) display the amplitude and phase line spectra of the M-periodic waveform, and parts (d) through (f) display the waveforms based on the first five terms, the first ten terms, and the first 1000 terms of the Fourier series, respectively.

t

(a) f1(t) f2(t) B

 As expected, the addition of more terms in the Fourier series improves the overall fidelity of the reproduced waveform. However, no matter how many terms are included in the series representation, the reproduction cannot duplicate the original M-waveform at points of discontinuity, such as when the waveform jumps from zero to 1. Discontinuities generate oscillations. 

0 T/2

−T/2

t

(b) f2(t) f3(t) A

Increasing the number of terms (adding more harmonics) reduces the period of the oscillation. Ultimately, the oscillations fuse into a solid line, except at the discontinuities (see expanded view of the discontinuity at t = −3 s in Fig. 13-6(f)).  As n approaches ∞, the Fourier series representation will reproduce the original waveform with perfect fidelity at all nondiscontinuous points, but at a point where the waveform jumps discontinuously between two different levels, the Fourier series will converge to a level half-way between them.  At t = 1 s, 3 s, 5 s, . . . , the Fourier series will converge to 0.5. This oscillatory behavior of the Fourier series in the neighborhood of discontinuous points is called the Gibbs phenomenon.

Example 13-4: Waveform Synthesis

Given that waveform f1 (t) in Fig. 13-7(a) is represented by the Fourier series f1 (t) =

∞  4A n=1



sin

 nπ  2

 cos

2nπ t T

 ,

0 T/2

−T/2

t

−A (c) f3(t) Figure 13-7: Waveforms for Example 13-4.

generate the Fourier series corresponding to the waveforms displayed in Figs. 13-7(b) and (c). Solution: (1) Waveform f2 (t) Waveforms f1 (t) and f2 (t) are similar in shape and have the same period, but they also exhibit two differences: (1) the dc value of f1 (t) is zero because it has dc symmetry, whereas the dc value of f2 (t) is B/2, and (2) the peak-to-peak value of f1 (t) is 2A, compared with only B for f2 (t). Mathematically, f2 (t) is related to f1 (t) by B f2 (t) = + 2



B 2A

 f1 (t)

  ∞  nπ  B  2B 2nπ t = + sin cos . 2 nπ 2 T n=1

“book” — 2015/5/4 — 7:27 — page 688 — #15

688

TECHNOLOGY BRIEF 30: BANDWIDTH, DATA RATE, AND COMMUNICATION

Technology Brief 30 Bandwidth, Data Rate, and Communication In Section 9-4.1, we defined the bandwidth B of a resonant circuit as the frequency span over which power transfer through the circuit is greater than half of the maximum level possible. This common half-power (or −3 dB) definition for B can be extended to many devices, circuits, and transmission channels. But how does the everyday use of the word bandwidth refer to the data rate of a transmission channel, such as the rate at which your internet connection can download data?

Signal and Noise in Communication Channels Every circuit (including switches, amplifiers, filters, phase shifters, rectifiers, etc.) and every transmission medium (air, wires, and optical fibers) operates with acceptable performance over some specific range of frequencies, outside of which ac signals are severely damped. The actual span of this operational frequency range is dictated by the physical characteristics of the circuit or transmission medium. One such example is the coaxial cable commonly used to connect a TV to a “cable network” or to an outside antenna. The coaxial cable is a high-fidelity transmission medium—causing negligible distortion or attenuation of the signal passing through it—so long as the carrier frequency of the signal is not much higher than about 10 GHz. The “cutoff” frequency of a typical coaxial cable is determined by the cable’s distributed capacitance, inductance, and resistance, which are governed in turn by the geometry of the cable, the conductivity of its inner and outer conductors, and the permittivity of the insulator that separates them. The MOSFET offers another example; in Section 5-7 we noted that the switching speed of a MOSFET circuit υin(t)

Peak-to-peak signal amplitude

υout(t) Noise amplitude

t

(a) Input

is limited by parasitic capacitances, setting an upper limit on the switching frequency that a given MOSFET circuit can handle. A circuit with a maximum switching speed of 100 ps, for example, cannot respond to frequencies greater than 1/100 ps (or 10 GHz) without distorting the output waveform in some significant way. Our third example is Earth’s atmosphere. According to Fig. TF20-1 (in Technology Brief 20: The Electromagnetic Spectrum), the transmission spectrum for the atmosphere is characterized by a limited set of transmission windows, with each window extending over a specific range of frequencies. The overall effective bandwidth B of a communication system is determined by the operational bandwidths of its constituent circuits and the transmission spectra of the cables or other transmission media it uses. As we will see shortly, the channel capacity (or data rate) of the system is directly proportional to B, but it also is influenced by the intensity and character of the noise in the system. Noise is random power self-generated by all real devices, circuits, and transmission media. In fact, any material at a temperature greater than 0 K (which includes all physical materials, since no material can exist at exactly 0 K) emits noise power all of the time. Figure TF30-1 illustrates how the noise generated by a circuit modifies the waveform of the signal passing through it. The input is an ideal sine wave, whereas the output consists of the same sine wave but with a fluctuating component added to it. The fluctuating component represents the noise generated by the circuit, which is random in polarity because the voltage associated with the noise fluctuates randomly between positive and negative values. If we know both the power PS carried by the signal and the average power PN associated with the noise, we then can determine the signal-to-noise ratio (SNR): PS SNR = . PN

t

Circuit

(b) Output

Figure TF30-1: The noise generated by the circuit adds a fluctuating component to the input signal.

“book” — 2015/5/4 — 7:27 — page 689 — #16

TECHNOLOGY BRIEF 30: BANDWIDTH, DATA RATE, AND COMMUNICATION

689

Table TT30-1: Theoretical limits for data transmission capacity given typical SNR and channel bandwidth number for several common digital communication layers.

Physical Channel Analog phone line 802.11g Wifi 100-BaseT Ethernet 10 Gb/s Ethernet

Channel Capacity (C) ~30 kbps 0

13-6.5

The Fourier transform of an exponentially decaying function that starts at t = 0 is F(ω) = F [Ae−at u(t)] ∞ =

Ae−at e−j ωt

0

which is problematic because e−j ∞ does not converge. To avoid the convergence problem, we can pursue an alternative approach that involves the signum function, which is defined by sgn(t) = u(t) − u(−t).

∞ e−(a+j ω)t  A . dt = A  = −(a + j ω)  a + jω 0

Hence, Ae−at u(t)

∞ e−j ωt  j = (e−j ∞ − 1),  −j ω 0 ω

A , a + jω

for a > 0.

(13.73)

(13.74)

Shown graphically in Fig. 13-17(a), the signum function resembles a step-function waveform (with an amplitude of 2 units) that has been slid downward by 1 unit. Looking at the waveform, it is easy to see that one can generate a step function from the signum function as follows: u(t) =

1 1 + sgn(t). 2 2

(13.75)

“book” — 2015/5/4 — 7:27 — page 707 — #34

13-6

FOURIER TRANSFORM PAIRS

707 Use of Eq. (13.78) in Eq. (13.76) gives

sgn(t)

F [u(t)] = π δ(ω) +

1 0

t

1 . jω

Equivalently, the preceding result can be expressed in the form

−1 (a) Signum function

u(t)

π δ(ω) +

sgn(t) e−εt u(t)

1 0 eεt u(−t)

t

(b) Model for sgn(t) Figure 13-17: The model shown in (b) approaches the exact definition of sgn(t) as ε → 0.

The corresponding Fourier transform is given by   1 1 1 F [u(t)] = F + F [sgn(t)] = π δ(ω) + F [sgn(t)], 2 2 2 (13.76) where in the first term we used the relationship given by Eq. (13.68b). Next, we will obtain F [sgn(t)] by modeling the signum function as sgn(t) = lim [e ε→0

εt

u(t) − e u(−t)]

(13.77)

with ε > 0. The shape of the modeled waveform is shown in Fig. 13-17(b) for a small value of ε. Now we are ready to apply the formal definition of the Fourier transform given by Eq. (13.62a): ∞ F [sgn(t)] = −∞

lim [e−εt u(t) − eεt u(−t)]e−j ωt dt

Example 13-10: Fourier Transform Properties

Establish the validity of the time derivative and modulation properties of the Fourier transform (Properties 6 and 10 in Table 13-5). Solution: Time derivative property: From Eq. (13.62b), 1 f (t) = 2π

= lim ⎣



ε→0



0

e−(ε+j ω)t dt −

0

−∞

⎤ e(ε−j ω)t dt ⎦

∞ 0 ⎤ −(ε+j ω)t  (ε−j ω)t  e e   ⎦ = lim ⎣  −  ε→0 −(ε + j ω)  ε − jω  0 −∞   1 1 2 = lim − = . (13.78) ε→0 ε + j ω ε − jω jω



F(ω) ej ωt dω.

(13.80)

−∞

Differentiating both sides with respect to t gives f (t) =

df 1 = dt 2π



j ω F(ω) ej ωt dω

−∞



ε→0



(13.79)

Table 13-4 provides a list of commonly used time functions together with their corresponding Fourier transforms, and Table 13-5 offers a summary of the major properties of the Fourier transform—many of which resemble those we encountered earlier in Chapter 12 in connection with the Laplace transform.

−1

−εt

1 . jω

1 = jω ⎣ 2π



⎤ F(ω) ej ωt dω⎦ .

−∞

Hence, differentiating f (t) in the time domain is equivalent to multiplying F(ω) by j ω in the frequency domain: f (t)

j ω F(ω).

(13.81)

“book” — 2015/5/4 — 7:27 — page 708 — #35

708

CHAPTER 13

FOURIER ANALYSIS TECHNIQUE

Table 13-4: Examples of Fourier transform pairs. Note that constant a ≥ 0. f (t)

F(ω) = F [f (t)]

|F(ω)|

BASIC FUNCTIONS 1.

δ(t)

1

δ(t)

1

1

t

ω

1 1a.

δ(t − t0 )

t

t0

ω

1

2.

1

e−j ωt0

1

2π δ(ω)

2π ω

t 1

3.

u(t)

π δ(ω) + 1/j ω

π

t

ω

1 t

4.

sgn(t)

2/j ω

ω

−1 τ 1 5.

τ

rect(t/τ )

t

|t|

6.

τ sinc(ωτ/2)

−2/ω2

ω

t 7.

e−at u(t)

1

1/a

1/(a + j ω)

t 8.

1

9.

1

ω

−2π 2π τ τ

ω

t

t

cos ω0 t

π[δ(ω − ω0 ) + δ(ω + ω0 )]

π −ω0

π ω0

sin ω0 t

j π[δ(ω + ω0 ) − δ(ω − ω0 )]

π −ω0

ω0 −π

ω

ω

ADDITIONAL FUNCTIONS 10. 11. 12. 13.

ej ω0 t

te−at u(t)

[e−at sin ω0 t] u(t) [e−at cos ω0 t] u(t)

Time modulation property: We start by multiplying both sides of Eq. (13.80) by cos ω0 t, and for convenience, we change the dummy variable ω to ω :

2π δ(ω − ω0 ) 1/(a + j ω)2 ω0 /[(a + j ω)2 + ω02 ] (a + j ω)/[(a + j ω)2 + ω02 ]

1 cos ω0 t f (t) = 2π

∞ −∞



cos ω0 t F(ω ) ej ω t dω .

“book” — 2015/5/4 — 7:27 — page 709 — #36

13-6

FOURIER TRANSFORM PAIRS

709

Table 13-5: Major properties of the Fourier transform. f (t)

Property

F(ω) = F [f (t)] K f (t)

1. Multiplication by a constant

K1 f1 (t) + K2 f2 (t)

2. Linearity

f (at)

3. Time scaling

f (t − t0 )

4. Time shift

ej ω0 t f (t)

5. Frequency shift

f =

6. Time 1st derivative

df dt

d nf dt n

7. Time nth derivative t 8. Time integral

e−j ωt0 F(ω) F(ω − ω0 ) j ω F(ω) (j ω)n F(ω) F(ω) + π F(0) δ(ω) jω

t n f (t)

(j )n

d n F(ω) dωn

cos ω0 t f (t)

1 2 [F(ω − ω0 ) + F(ω + ω0 )]

11. Convolution in t

f1 (t) ∗ f2 (t)

F1 (ω) F2 (ω)

f1 (t) f2 (t)

Applying Euler’s identity to cos ω0 t on the right-hand side leads to

1 = 4π

1 ω F |a| a

10. Modulation

12. Convolution in ω

1 cos ω0 t f (t) = 2π

K1 F1 (ω) + K2 F2 (ω)

f (t) dt −∞

9. Frequency derivative

K F(ω)

∞  −∞

 ∞

ej ω0 t + e−j ω0 t 2



1 F1 (ω) ∗ F2 (ω) 2π

the second integral, we have  ∞ 1 1 F(ω − ω0 ) ej ωt dω cos ω0 t f (t) = 2 2π −∞



F(ω ) ej ω t dω +



F(ω ) ej (ω +ω0 )t dω

1 2π



 F(ω + ω0 ) ej ωt dω ,

−∞

which can be cast in the abbreviated form

−∞

∞ +



F(ω ) e

j (ω −ω0 )t





dω .

cos ω0 t f (t)

1 [F(ω − ω0 ) + F(ω + ω0 )]. 2 (13.82)

−∞

Upon making the substitution (ω = ω +ω0 ) in the first integral and independently making the substitution (ω = ω − ω0 ) in

Concept Question 13-14: What is the Fourier transform

of a dc voltage? (See

)

“book” — 2015/5/4 — 7:27 — page 710 — #37

710

CHAPTER 13

Concept Question 13-15: “An impulse in the time

Exercise 13-9), which is given by

domain is equivalent to an infinite number of sinusoids, all with equal amplitude.” Is this a true statement? Can one construct an ideal impulse function? (See ) Exercise 13-7: Use the entries in Table 13-4 to determine the Fourier transform of u(−t). Answer: F(ω) = π δ(ω) − 1/j ω. (See

FOURIER ANALYSIS TECHNIQUE

F(−ω) = F∗ (ω). (reversal property)

The combination of Eqs. (13.83) and (13.84) can be written as

)



1 f (t) dt = 2π

∞ |F(ω)|2 dω.

2

Exercise 13-8: Verify the Fourier transform expression

−∞

for entry #10 in Table 13-4. Answer: (See

13-6.7



∞ f 2 (t) dt =

 Parseval’s theorem states that the total energy in the time domain is equal to the total energy in the ω domain. 

⎡ f (t) ⎣

−∞

1 2π



⎤ F(ω) ej ωt dω⎦ dt,

−∞

(13.83) where one f (t) was replaced with the inverse Fourier transform relationship given by Eq. (13.62b). By reversing the order of f (t) and F(ω), and reversing the order of integration, we have 1 W = 2π 1 = 2π =

1 2π

1 = 2π

∞ −∞

∞ −∞

⎡ F(ω) ⎣



⎤ f (t) ej ωt dt ⎦ dω

−∞

⎡ F(ω) ⎣



⎤ f (t) e−j (−ω)t dt ⎦ dω

−∞



Exercise 13-9: Verify the reversal property given by

Eq. (13.85). Answer: (See

13-7

)

Phasor vs. Laplace vs. Fourier

Consider a linear circuit with input excitation x(t) and output response y(t), where x(t) is a voltage or current source and y(t) is a voltage between two nodes in the circuit or a current through one of its branches. Our goal is to analyze the circuit to determine the desired output y(t). Beyond the time-domain differential equation solution method, which in practice can accommodate only first- and second-order circuits, we have available to us three techniques by which to determine y(t). (a) The phasor-domain technique (Chapter 7). (b) The Laplace transform technique (Chapter 12). (c) The Fourier series and transform techniques (Chapter 13).

F(ω) F(−ω) dω −∞



−∞

(Parseval’s theorem)

Parseval’s Theorem

−∞

(13.86)

)

If f (t) represents the voltage across a 1  resistor, then f 2 (t) represents the power dissipated in the resistor, and the integrated value of f 2 (t) over [−∞, ∞] represents the cumulative energy W expended in the resistor. Thus,

W =

(13.85)

F(ω) F∗ (ω) dω,

(13.84)

−∞

where we used the reversal property (also known as the conjugate symmetry property) of the Fourier transform (see

The applicability conditions for the three techniques, summarized in Table 13-6, are governed by the duration and shape of the waveform of the input excitation. Based on its duration, an input excitation x(t) is said to be: (1) everlasting (two-sided): if it exists over all time (−∞, ∞), (2) one-sided: if it starts at or after t = 0; i.e., x(t) = 0 for t < 0− .

“book” — 2015/5/4 — 7:27 — page 711 — #38

13-8

CIRCUIT ANALYSIS WITH FOURIER TRANSFORM

In real life, there is no such thing as an everlasting excitation. When we deal with real circuits, there is always a starting point in time for both the input excitation and output response. In general, an output signal consists of two components, a transient component associated with the initial onset of the input signal, and a steady-state component that alone remains after the transient component decays to zero. If the input excitation is sinusoidal and we are interested in only the steadystate component of the output response, it is often convenient to regard the input excitation as everlasting, even though, strictly speaking, it cannot be so. We regard it as such because we can then apply the phasor-domain technique, which is easier to implement than the other two techniques. According to the summary provided in Table 13-6:

711

R1 υs(t)

• If x(t) is a one-sided excitation, the Laplace transform technique is the preferred solution method. An important feature of the technique is that it can accommodate nonzero initial conditions of the circuit, if they exist. • If x(t) is everlasting and its waveform is nonperiodic, we can obtain y(t) by applying either a bilateral form of the Laplace transform or the Fourier transform. For input excitations x(t) whose Laplace transforms do not exist but their Fourier transforms do, the Fourier transform approach becomes the only viable option, and the converse is true for excitations whose Fourier transforms do not exist but their Laplace transforms do. The Laplace transform operates in the s domain, where s = σ + j ω is represented by a complex plane with real and imaginary axes along σ and j ω. The Fourier transform operates along the j ω axis of the splane, corresponding to σ = 0. Thus, for circuits excited by everlasting sinusoidal waveforms,  F(ω) = F(s)σ =0 .

R2

C

(a) Time domain R1 IC

+ Vs(ω) _

R2

1 jωC

(b) ω-domain

• If x(t) is an everlasting sinusoid, the phasor-domain technique is the solution method of choice. • If x(t) is an everlasting periodic excitation, such as a square wave or any repetitive waveform that can be represented by a Fourier series, then by virtue of the superposition principle, the phasor-domain technique can be used to compute the output responses corresponding to the individual Fourier components of the input excitation, and then all of the output components can be added up to generate the total output.

iC(t)

+ _

Figure 13-18: Circuits for Example 13-11.

13-8

Circuit Analysis with Fourier Transform

As was mentioned earlier, the Fourier transform technique can be used to analyze circuits excited by either one-sided or two-sided nonperiodic waveforms—as long as the circuit has no initial conditions. The procedure (which is analogous to the Laplace transform technique) with s replaced by j ω is demonstrated through Example 13-11. Example 13-11: RC Circuit

The RC circuit shown in Fig. 13-18(a) is excited by a voltage source υs (t). Apply Fourier analysis to determine iC (t) if: (a) υs = 10u(t), (b) υs (t) = 10e−2t u(t), and (c) υs (t) = 10+5 cos 4t, all measured in volts. The element values are R1 = 2 k, R2 = 4 k, and C = 0.25 mF. Solution: Step 1: Transfer circuit to ω domain In the frequency-domain circuit shown in Fig. 13-18(b), Vs (ω) is the Fourier transform of υs (t). Step 2: Determine H(ω) = IC (ω)/Vs (ω) Application of source transformation to the circuit in Fig. 13-18(b) followed with current division leads to H(ω) =

IC (ω) j ω/R1 j 0.5ω × 10−3 = . = R1 + R 2 Vs (ω) 3 + jω + jω R1 R2 C (13.87)

“book” — 2015/5/4 — 7:27 — page 712 — #39

712

CHAPTER 13

FOURIER ANALYSIS TECHNIQUE

Table 13-6: Methods of solution. Input x(t) Duration

Waveform

Solution Method

Output y(t)

Everlasting

Sinusoid

Phasor Domain

Steady-State Component (no transient exists)

Everlasting

Periodic

Phasor Domain and Fourier Series

Steady-State Component (no transient exists)

One-sided, x(t) = 0, for t < 0−

Any

Laplace Transform (unilateral) (can accommodate nonzero initial conditions)

Complete Solution (transient + steady-state)

Everlasting

Any

Bilateral Laplace Transform or Fourier Transform

Complete Solution (transient + steady-state)

Step 3: Solve for IC (ω) and iC (t)

Application of partial fraction expansion (Section 12-5) gives

(a) υs (t) = 10u(t): The corresponding Fourier transform per entry #3 in Table 13-4 is Vs (ω) = 10π δ(ω) +

IC (ω) = with

10 . jω

A1 = (2 + j ω) IC (ω)|j ω=−2  j 5ω × 10−3  = = −10 × 10−3 3 + j ω j ω=−2

The corresponding current is j 5π ω δ(ω) × 10−3 5 × 10−3 + . 3 + jω 3 + jω (13.88) The inverse Fourier transform of IC (ω) is given by IC (ω) = H(ω) Vs (ω) =

1 iC (t) = 2π

∞ −∞

+F

−1

j 5πω δ(ω) × 10−3 j ωt e dω 3 + jω 

−3 

5 × 10 3 + jω

(b) υs (t) =

(13.89)



× 10−3

and iC (t) = (15e−3t − 10e−2t ) u(t) mA.

(13.91)

(c) υs (t) = 10 + 5 cos 4t: By entries #2 and #8 in Table 13-4, Vs (ω) = 20π δ(ω) + 5π[δ(ω − 4) + δ(ω + 4)], IC (ω) = H(ω) Vs (ω)

10 . 2 + jω

j 5ω × 10−3 . (2 + j ω)(3 + j ω)

15 −10 + 2 + jω 3 + jω

and the capacitor current is

=

The corresponding current IC (ω) is given by IC (ω) = H(ω) Vs (ω) =

 IC (ω) =

u(t): By entry #7 in Table 13-4, Vs (ω) =

A2 = (3 + j ω) IC (ω)|j ω=−3  j 5ω × 10−3  = = 15 × 10−3 . 2 + j ω j ω=−3 Hence,

where we applied the formal definition of the inverse Fourier transform to the first term—because it includes a delta function—and the functional form to the second term—because we intend to use look-up entry #7 in Table 13-4. Accordingly,

10e−2t

and

,

iC (t) = 0 + 5e−3t u(t) mA.

A2 A1 + , 2 + jω 3 + jω

(13.90)

j 10π ω δ(ω) × 10−3 3 + jω   ω δ(ω + 4) −3 ω δ(ω − 4) + j 2.5π × 10 + . 3 + jω 3 + jω

“book” — 2015/5/4 — 7:27 — page 713 — #40

13-9

MULTISIM: MIXED-SIGNAL CIRCUITS AND THE SIGMA-DELTA MODULATOR

The corresponding time-domain current is obtained by applying Eq. (13.62b) as 1 iC (t) = 2π

∞ −∞

1 + 2π 1 + 2π

j 10πω δ(ω) × 10−3 ej ωt dω 3 + jω ∞ −∞

∞ −∞

j 2.5π ω × 10−3 δ(ω − 4) ej ωt dω 3 + jω j 2.5π ω × 10−3 δ(ω + 4) ej ωt dω 3 + jω

j 5 × 10−3 e−j 4t j 5 × 10−3 ej 4t − 3 + j4 3 − j4 $ % ◦ ◦ j 4t ej 36.9 e−j 4t e−j 36.9 −3 e = 5 × 10 + 5 5 =0+

= 2 cos(4t + 36.9◦ ) mA.

(13.92)

13-9 Multisim: Mixed-Signal Circuits and the Sigma-Delta Modulator Historically, circuit designers tended to fall into two broad classes: those who designed digital circuits and those who designed analog circuits. As a broad generalization, digitalcircuit designers built logic gates, computational elements, memories, and so on, whereas analog designers tended to work on circuits that interfaced with the noncircuit world: amplifiers, drivers, radio frequency circuits, analog-to-digital converters, among others. Moreover, digital designers tended to have more comprehensive and powerful software design tools, mainly because digital circuits could be abstracted into modules that made hierarchical analysis possible. For example, a transistor could be modeled as a simple switch, several of these switches could be wired together as a simple logic gate, many logic gates could be wired together to make a counter or a memory, and so on, all of which can be readily modeled as “black boxes” in software. Analog circuits, by contrast, defied this type of compartmentalization due to feedback loops, nonlinear behavior, and complex topologies; this made analog design almost an art form. Advances in silicon fabrication technologies have now blurred the line between these two worlds considerably. A new generation of circuits, known as mixed-signal circuits, contains elements of both worlds (Fig. 13-19). This exciting area combines the power of analog designs with the scalability,

713

modularity and computational power of digital circuits. Modern analog-to-digital conversion (ADC) and digital-to-analog conversion (DAC) circuits, cell phone communication circuits, software radio, internet routers, and audio synthesis circuits are all examples of mixed signal circuits. The advantages are numerous. Consider software radio, for example. We saw in Section 9-8 how a superheterodyne receiver works, which is a perfect example of a multistage analog circuit. But what if many of the functions of the superheterodyne receiver could be performed by digital circuits instead? What advantages might there be? One obvious advantage is the introduction of computational “intelligence” into the radio itself. If the receiver is designed, in part or in whole, with digital circuits, these circuits can be built around computation and memory. Programs can be loaded that allow the radio to change its power consumption, transmission patterns and protocols based on user or environmental parameters; this is often known as cognitive radio. The integration of analog and digital circuits comes with certain drawbacks, however. Design remains a challenging, and highly paid, exercise. Design and testing software for mixed signal circuits is nowhere near as advanced as that for digital circuits. The fabrication of these circuits is often confined to specialized processes not compatible with standard, digitalprocessor fabrication methodologies (although this is rapidly changing).

13-9.1 The Sigma-Delta ( ) Modulator and Analog-to-Digital Converters So far in this book we have used Multisim to model amplifiers, digital circuits, filters, resonators, and circuits that employ feedback. In this section, we will put it all together and show you how to design a very useful circuit, the Sigma-Delta ( ) Modulator, which since its early development by Inose and Yasuda in 1962 has now become a standard tool for the design of inexpensive analog-to-digital converters (ADCs). There are many ways to convert an analog waveform into a digital sequence of pulses. The classic ADC circuit takes a timevarying analog voltage, υin (t), and produces a corresponding time-varying digital output consisting of a number of bits (Vout0 , Vout1 , etc.). Figure 13-20(a) shows the process schematically. Here, a linearly increasing voltage is fed into a 4-bit ADC; as the input voltage changes with time, the four digital output bits change their state (either “0” or ”1”). All of the pulses have the same duration and can change states instantaneously. With 4 binary bits, we can construct 16 different values (i.e., 0000, 0001, . . . , 1111), so this 4-bit ADC converts any input voltage to one of 24 or, equivalently, 16 different digital values. Modern ADCs commonly have 12, 16 or even 24 output bits, giving them very high resolution (e.g., 224 = 16,777,216 different values!).

“book” — 2015/5/4 — 7:27 — page 714 — #41

714

CHAPTER 13

FOURIER ANALYSIS TECHNIQUE

Figure 13-19: This mixed-signal chip implements a highly reconfigurable RF receiver based on a down-converting Sigma-Delta A/D (courtesy Renaldi Winoto and Prof. Borivoje Nikolic, U.C. Berkeley)

Voltage

υout3 υin

υout2

(a) Traditional ADC

υout0

Voltage

Time

υout1

Time

υin Time

(b) Σ∆ ADC

υout Time

Figure 13-20: (a) A traditional 4-bit ADC converts an analog input voltage and produces 4 digital output bits; (b) a ADC generates a pulse train where the pulse duration is governed by the magnitude of the input voltage.

One usually trades off speed for resolution; the bits in a fast 12-bit ADC integrated circuit can change states once every 2

microseconds, which means that theADC can measure the input voltage approximately 500,000 times per second. Unlike conventional ADCs, the ADC generates an output consisting of a single digital bit. The duration of the voltage pulse, however, depends on the value of the input voltage (Fig. 13-20(b)), thereby encoding the magnitude of the input voltage into the duration of a single pulse, instead of encoding it into the binary states of several pulses (bits) of equal duration. The modulator is particularly attractive for designing and building inexpensive ADC circuits because it can (a) be made using digital components, which are less expensive to build and easier to test than analog components, and (b) the digital components can be reprogrammed and modified by the user using firmware. Although invented in the 1960s, the

modulator was not used commercially until digital CMOS processes became sufficiently fast (to produce time-varying output pulses faster than the changes exhibited by the input signal), small (enabling the mixing of digital with analog circuit components) and inexpensive to fabricate. The entire circuit (Fig. 13-21) can be built from analog components introduced in this book, namely a subtractor,

“book” — 2015/5/4 — 7:27 — page 715 — #42

13-9

MULTISIM: MIXED-SIGNAL CIRCUITS AND THE SIGMA-DELTA MODULATOR

υin(t) Analog signal

Difference amp

+ _

Integrator amp

Comparator amp

dt

715

Low pass filter υout(t) Digital pulse stream

1−bit DAC Figure 13-21: Block diagram of a modulator.

an integrator, a comparator, a 1-bit DAC and a low-pass filter. In real implementations, most of these components are replaced with digital substitutes, often reprogrammable during operation. Thus, one could replace the analog filter with an adjustable digital filter, the integrator with a digital integrator, and so on. The only analog component in modern s is usually the DAC.

13-9.2

How the Works

Table 13-6 shows the individual subcircuits of the modulator and Fig. 13-22 displays the complete circuit, all drawn in Multisim. Our basic circuit takes an analog input, υin (t), subtracts from it a feedback signal, υbit (t), then integrates this signal, producing υint (t). The integrated signal is then compared with a reference voltage (in our case, 0 V)

Figure 13-22: Complete Multisim circuit of the modulator.

“book” — 2015/5/4 — 7:27 — page 716 — #43

716

CHAPTER 13

FOURIER ANALYSIS TECHNIQUE

Table 13-7: Multisim circuits of the modulator. Multisim Circuit

Description and Notes

Subtractor: This is a difference amplifier (following Table 4-3) with a voltage gain of 1. VPLUS and VMINUS are the extremes of the analog input (in the complete circuit, they are set to ±12 V).

Vbit

Integrator: This circuit consists of an inverting integrator amplifier (Section 5-6.1) and an inverting amplifier (following Table 4-3) with a voltage gain of 1 (to remove the integrator’s negative sign).

Comparator: The comparator is a simple op amp with no feedback (open loop). Since the internal voltage gain A of the op amp is so high (Section 4-1.2), any positive difference between the noninverting and the inverting inputs immediately drives the amplifier output to VDD ; a negative difference drives the amplifier output to 0 V. VDD is set to the desired digital voltage level (5 V, in the case of the complete circuit in Fig. 13-22).

vout

1-Bit Digital-to-Analog Converter (DAC): The DAC is very similar to the comparator. The input voltage is compared to a voltage level halfway between 0 and VDD ; this has the effect of transforming an input of VDD into an output voltage of VPLUS/2 (+6 V in Fig. 13-22) and an input voltage of 0 V into an output voltage of VMINUS/2 (−6 V in Fig. 13-22).

and produces υout (t). The output of the comparator, υout (t), can only have two values: VDD or 0, where VDD is the dc power supply voltage of the comparator. Hence, υout (t) is a

time-varying digital signal. Note that υout (t) is also sent to a 1-bit DAC that converts the digital signal to an analog signal, υbit (t), which is fed back to the subtractor.

“book” — 2015/5/4 — 7:27 — page 717 — #44

13-9

MULTISIM: MIXED-SIGNAL CIRCUITS AND THE SIGMA-DELTA MODULATOR

The overall functionality of the modulator is illustrated by Fig. 13-23 for an input signal composed of a 1 Hz sinusoid with an amplitude of 4 V. We observe that the corresponding output, υout (t), consists of a sequence of pulses whose durations are proportional to the instantaneous level of the input voltage, υin (t). Thus, the circuit encodes amplitude information contained in an analog signal into pulse-duration information in a digital sequence. After transmission of υout (t) through downstream digital circuits, the original information can be retrieved by measuring the durations of the pulses. This can be accomplished by a digital counter, either in hardware (using a counter circuit) or in software on a microcontroller. In hardware, transitions can be detected with Schmidt triggers or similar edge detectors, and the counter is made to “count” the duration between transitions.

717

Figure 13-23: A 1 Hz sinusoidal ac signal, υin (t), blue trace, is converted to a series of pulses at the output, υout (t), red trace, by the Sigma Delta modulator. Note that the duration of the pulses is related to the instantaneous level of voltage υin (t).

Summary Concepts • A periodic waveform of period T can be represented by a Fourier series consisting of a dc term and sinusoidal terms that are harmonic multiples of ω0 = 2π/T . • The Fourier series can be represented in terms of a cosine/sine form, amplitude/phase form, and a complex exponential form. • Circuits excited by a periodic waveform can be analyzed by applying the superposition theorem to the individual terms of the harmonic series.

• Nonperiodic waveforms can be represented by a Fourier transform. • Upon transforming the circuit to the frequency domain, the circuit can be analyzed for the desired voltage or current of interest and then the result can be inverse transformed to the time domain. • The Sigma-Delta modulator is an example of a mixedsignal circuit. It converts an analog waveform into a single-bit digital pulse whose duration is proportional to the instantaneous magnitude of the waveform.

Mathematical and Physical Models Fourier Series

Table 13-3

sinc(x) =

sinc Function

sin x x

Average Power ∞

Pav = Vdc Idc +

1 Vn In cos(φυn − φin ) 2 n=1

Fourier Transform ∞ F(ω) = F [f (t)] =

f (t) e−j ωt dt

−∞

f (t) = F −1 [F(ω)] =

1 2π

∞ −∞

F(ω) ej ωt dω

Properties of Fourier Transform

Table 13-5

2-D Fourier Transform F(ωx , ωy ) = F [f (x, y)] ∞ = f (x, y) e−j ωx x e−j ωy y dx dy −∞

1 (2π )2 ∞ = F(ωx , ωy ) ej ωx x ej ωy y dωx dωy

f (x, y) =

−∞

“book” — 2015/5/4 — 7:27 — page 718 — #45

718

CHAPTER 13

Important Terms

FOURIER ANALYSIS TECHNIQUE

Provide definitions or explain the meaning of the following terms:

ADC absolutely integrable amplitude of the nth harmonic amplitude/phase amplitude spectrum cognitive radio complete conjugate symmetry property dc component Dirichlet conditions duration even function even symmetry

everlasting everlasting periodic excitation everlasting sinusoid exponential form Fourier coefficient Fourier series Fourier theorem Fourier transform frequency-shift property frequency spectrum fundamental angular frequency Gibbs phenomenon

harmonic inverted line spectrum magnitude spectrum mixed signal circuit nmax -truncated series null odd function odd symmetry one-sided one-sided excitation Parseval’s theorem periodic function periodicity property

PROBLEMS

periodic waveform phase phase spectrum reversal property Schmidt triggers Sigma-Delta modulator signum function sinc function spacing between adjacent harmonics steady-state component time-shift property transient component truncated series

f(t) A

Sections 13-1 and 13-2: Fourier Series

0

For each of the waveforms in Problems 13.1 through 13.10: (a) Determine if the waveform has dc, even, or odd symmetry.

−8

(b) Obtain its cosine/sine Fourier series representation.

−6

−4

−2

t (s) 0

2

4

6

8

Figure P13.2: Waveform for Problem 13.2.

(c) Convert the representation to amplitude/phase format and plot the line spectra for the first five nonzero terms. (d) Use MATLAB software to plot the waveform using a truncated Fourier series representation with nmax = 100.

13.3 Waveform in Fig. P13.3 with A = 6.

13.1 Waveform in Fig. P13.1 with A = 10.

f(t) f(t)

A

A 0 −4

0

4

8

t (s)

−3

−2

−1

t (s) 0

1

2

3

Figure P13.3: Waveform for Problem 13.3. Figure P13.1: Waveform for Problem 13.1.

13.2 Waveform in Fig. P13.2 with A = 4.

13.4 Waveform in Fig. P13.4 with A = 10.

“book” — 2015/5/4 — 7:27 — page 719 — #46

PROBLEMS

719 13.8 Waveform in Fig. P13.8 with A = 10.

f(t) A

f(t)

0

A

t (s)

−6 −5 −4 −3 −2 −1 0 1 2 3 4 5 6 −4

−3

−1

−2

Figure P13.4: Waveform for Problem 13.4.

t (ms)

0

2

1

3

4

−A

*13.5 Waveform in Fig. P13.5 with A = 20.

Figure P13.8: Waveform for Problem 13.8.

f(t)

*13.9 Waveform in Fig. P13.9 with A = 10.

A

f(t)

0

t (s)

−6 −5 −4 −3 −2 −1

A

0 1 2 3 4 5 6 7

A/2

−A −2

Figure P13.5: Waveform for Problem 13.5.

−1

0

1

2

3

4

t (s)

−A/2 13.6 Waveform in Fig. P13.6 with A = 100.

−A Figure P13.9: Waveform for Problem 13.9.

f(t) A

13.10 Waveform in Fig. P13.10 with A = 20.

0 −6

−4

−2

t (s) 0

2

4

6

f(t) A

Figure P13.6: Waveform for Problem 13.6.

−4

13.7 Waveform in Fig. P13.7 with A = 4.

−3

−2

−1 0 −A

1

2

3

4

t (s)

f(t) Figure P13.10: Waveform for Problem 13.10.

−8

−6

−4

−2

0

2

4

6

8

−A Figure P13.7: Waveform for Problem 13.7.



Answer(s) available in Appendix G.

t (s) 13.11 Obtain the cosine/sine Fourier series representation for f (t) = cos2 (4π t), and use MATLAB software to plot it with nmax = 2, 10, and 100. *13.12

Repeat Problem 13.11 for f (t) = sin2 (4π t).

13.13

Repeat Problem 13.11 for f (t) = | sin(4π t)|.

“book” — 2015/5/4 — 7:27 — page 720 — #47

720

CHAPTER 13

13.14 Which of the six waveforms shown in Figs. P13.1 through P13.6 will exhibit the Gibbs oscillation phenomenon when represented by a Fourier series? Why? 13.15 Consider the sawtooth waveform shown in Fig. 13-3(a). Evaluate the Gibbs phenomenon in the neighborhood of t = 4 s by plotting the Fourier series representation with nmax = 100 over the range between 3.99 s and 4.01 s (using expanded scales if necessary). *13.16 The Fourier series of the periodic waveform shown in Fig. P13.16(a) is given by f1 (t) = 10 −

20 π

∞  n=1



1 nπ t sin n 2

 .

Determine the Fourier series of waveform f2 (t) in Fig. P13.16(b).

FOURIER ANALYSIS TECHNIQUE

(c) Plot υout (t) using nmax = 100.

R1 + υs(t) _

R2

+ υout(t) _

C

Figure P13.17: Circuit for Problem 13.17.

13.18 The current source is (t) in the circuit of Fig. P13.18 generates a sawtooth wave (waveform in Fig. 13-3(a)) with a peak amplitude of 20 mA and a period T = 5 ms. (a) Derive the Fourier series representation of υout (t). (b) Calculate the first five terms of υout (t) using R1 = 500 , R2 = 2 k, and C = 0.33 μF. (c) Plot υout (t) and is (t) using nmax = 100.

f1(t) i2

20

−4

R2

+

i1 0

4

8

t (s)

is(t)

υout(t)

C

R1

_

(a) f1(t) Figure P13.18: Circuit for Problem 13.18.

f2(t) 20 0 −4

0

4

8

t (s)

−20 (b) f2(t)

13.19 The current source is (t) in the circuit of Fig. P13.19 generates a train of pulses (waveform #3 in Table 13-2) with A = 6 mA, τ = 1 μs, and T = 10 μs. (a) Derive the Fourier series representation of i(t). (b) Calculate the first five terms of i(t) using R = 1 k, L = 1 mH, and C = 1 μF. (c) Plot i(t) and is (t) using nmax = 100.

i(t)

Figure P13.16: Waveforms for Problem 13.16.

Section 13-3: Circuit Applications 13.17 The voltage source υs (t) in the circuit of Fig. P13.17 generates a square wave (waveform #1 in Table 13-2) with A = 10 V and T = 1 ms. (a) Derive the Fourier series representation of υout (t). (b) Calculate the first five terms of υout (t) using R1 = R2 = 2 k and C = 1 μF.

is(t)

R

L C

Figure P13.19: Circuit for Problem 13.19.

13.20 Voltage source υs (t) in the circuit of Fig. P13.20(a) has the waveform displayed in Fig. P13.20(b).

“book” — 2015/5/4 — 7:27 — page 721 — #48

PROBLEMS

721

(a) Derive the Fourier series representation of i(t). (b) Calculate the first five terms of i(t) using R1 = R2 = 10  and L1 = L2 = 10 mH. (c) Plot i(t) and υs (t) using nmax = 100.

(c) Plot υout (t) and υs (t) using nmax = 100. 13.23 Determine υout (t) in the circuit of Fig. P13.23, given that the input excitation is characterized by a triangular waveform (#4 in Table 13-2) with A = 24 V and T = 20 ms. (a) Derive Fourier series representation of υout (t).

R2

R1 + υs(t) _

L1

i(t)

(b) Calculate first five terms of υout (t) using R = 470 , L = 10 mH, and C = 10 μF.

L2

(c) Plot υout (t) and υs (t) using nmax = 100.

(a) Circuit

L

υs(t)

+ υs(t) _

10 V

C

+ υout(t) _

Figure P13.23: Circuit for Problem 13.23.

2V −1

R

0

1

2

t (ms)

(b) Waveform Figure P13.20: Circuit and waveform for Problem 13.20.

13.24 A backward-sawtooth waveform (#7 in Table 13-2) with A = 100 V and T = 1 ms is used to excite the circuit in Fig. P13.24. (a) Derive Fourier series representation of υout (t).

13.21 Determine the output voltage υout (t) in the circuit of Fig. P13.21, given that the input voltage υin (t) is a full-wave rectified sinusoid (waveform #8 in Table 13-2) with A = 120 V and T = 1 μs. (a) Derive the Fourier series representation of υout (t). (b) Calculate the first five terms of υout (t) using R = 1 k, L = 1 mH, and C = 1 nF. (c) Plot υout (t) and υin (t) using nmax = 100.

(c) Plot υout (t) and υs (t) using nmax = 100.

R1 + υs(t) _

+

R2 C L

L + υin(t) _

(b) Calculate the first five terms of υout (t) using R1 = 1 k, R2 = 100 , L = 1 mH, and C = 1 μF.

C

R

+ υout(t) _

Figure P13.21: Circuit for Problem 13.21.

13.22 (a) Repeat Example 13-5, after replacing the capacitor with an inductor L = 0.1 H and reducing the value of R to 1 . (b) Calculate the first five terms of υout (t).

υout(t)

_

Figure P13.24: Circuit for Problem 13.24.

13.25 The circuit in Fig. P13.25 is excited by the source waveform shown in Fig. P13.20(b). (a) Derive Fourier series representation of i(t). (b) Calculate the first five terms of υout (t) using R1 = R2 = 100 , L = 1 mH, and C = 1 μF. (c) Plot i(t) and υs (t) using nmax = 100.

“book” — 2015/5/4 — 7:27 — page 722 — #49

722

CHAPTER 13

R1

(a) Determine the exact value of the average power consumed by the resistor.

R2 i(t)

+ υs(t) _

C

(b) Using a truncated Fourier series representation of the waveform with only the first four terms, obtain an approximate value for the average power consumed by the resistor.

L

(c) What is the percentage of error in the value given in (b)?

Figure P13.25: Circuit for Problem 13.25.

*13.26 The RC op-amp integrator circuit of Fig. P13.26 is excited by a square wave (waveform #1 in Table 13-2) with A = 4 V and T = 2 s. (a) Derive Fourier series representation of υout (t). (b) Calculate the first five terms of υout (t) using R1 = 1 k, R1 = 10 k, and C = 10 μF. (c) Plot υout (t) using nmax = 100.

_

+ υs(t) _

+

13.31 The current source in the parallel RLC circuit of Fig. P13.31 is given by is (t) = [10 + 5 cos(100t + 30◦ ) − cos(200t − 30◦ )] mA. Determine the average power dissipated in the resistor given that R = 1 k, L = 1 H, and C = 1 μF.

is(t)

C R1

FOURIER ANALYSIS TECHNIQUE

R2

+

L

C

R

Figure P13.31: Circuit for Problem 13.31.

υout(t)

Figure P13.26: Circuit for Problem 13.26.

13.27 Repeat Problem 13.26 after interchanging the locations of the 1 k resistor and the 10 μF capacitor.

13.32 A series RC circuit is connected to a voltage source whose waveform is given by waveform #5 in Table 13-2, with A = 12 V and T = 1 ms. Using a truncated Fourier series representation composed of only the first three nonzero terms, determine the average power dissipated in the resistor, given that R = 2 k and C = 1 μF.

Section 13-4: Average Power

Sections 13-5 and 13-6: Fourier Transform

13.28 The voltage across the terminals of a certain circuit and the current entering into its (+) voltage terminal are given by

For each of the waveforms in Problems 13.33 through 13.42, determine the Fourier transform.

υ(t) = [4 + 12 cos(377t + 60◦ ) − 6 cos(754t − 30◦ )] V, ◦

*13.33 Waveform in Fig. P13.33 with A = 5 and T = 3 s.



i(t) = [5 + 10 cos(377t + 45 ) + 2 cos(754t + 15 )] mA. Determine the average power consumed by the circuit, and the ac power fraction. *13.29 The current flowing through a 2 k resistor is given by

f(t) A

i(t) = [5 + 2 cos(400t + 30◦ ) + 0.5 cos(800t − 45◦ )] mA. Determine the average power consumed by the resistor, as well as the ac power fraction. 13.30 The current flowing through a 10 k resistor is given by a triangular waveform (#4 in Table 13-2) with A = 4 mA and T = 0.2 s.

T

t (s)

Figure P13.33: Waveform for Problem 13.33.

13.34 Waveform in Fig. P13.34 with A = 10 and T = 6 s.

“book” — 2015/5/4 — 7:27 — page 723 — #50

PROBLEMS

723

f(t)

f(t) A

A

T/2

−T/2

T/3 2T/3

0

t (s)

Figure P13.34: Waveform for Problem 13.34.

T

t (s)

−A Figure P13.37: Waveform for Problem 13.37.

13.35 Waveform in Fig. P13.35 with A = 12 and T = 3 s. 13.38 Waveform in Fig. P13.38 with A = 1 and T = 2 s.

f(t) f(t) A A 2A/3 A/2

A/3 T/3

2T/3

T

t (s) t (s)

T/2

−T/2 Figure P13.35: Waveform for Problem 13.35.

Figure P13.38: Waveform for Problem 13.38.

13.36 Waveform in Fig. P13.36 with A = 2 and T = 12 s.

*13.39 Waveform in Fig. P13.39 with A = 3 and T = 1 s.

f(t)

f(t)

A

A −2T

−T

A/2

T

T/3

2T/3

T

t (s)

Figure P13.36: Waveform for Problem 13.36.

13.37 Waveform in Fig. P13.37 with A = 1 and T = 3 s.

2T

t (s)

−A Figure P13.39: Waveform for Problem 13.39.

13.40 Waveform in Fig. P13.40 with A = 5, T = 1 s, and α = 10 s−1 .

“book” — 2015/5/4 — 7:27 — page 724 — #51

724

CHAPTER 13 (a) F (3t − 2)

f(t) A(1 −

FOURIER ANALYSIS TECHNIQUE

*(b) t f (t) (c) d f (t)/dt

e−αt )

13.47

Let the Fourier transform of f (t) be

t (s)

T

F (ω) =

Figure P13.40: Waveform for Problem 13.40.

13.41 Waveform in Fig. P13.41 with A = 10 and T = 2 s.

f(t)

1 e−j ω + B. (A + j ω)

Determine the Fourier transforms of the following signals (set A = 2 and B = 1).   (a) f 58 t (b) f (t) cos(At) (c) d 3 f/dt 3

A A cos(2πt/T)

13.48

Prove the following two Fourier transform pairs.

(a) cos(ωT ) F (ω) (b) sin(ωT ) F (ω)

T/4

−T/4

t (s)

1 2 [f (t − T ) + f (t + T )] 1 2j [f (t + T ) − f (t − T )]

Section 13-8: Circuit Analysis with Fourier Transform

Figure P13.41: Waveform for Problem 13.41.

13.49 The circuit in Fig. P13.19 is excited by the source waveform shown in Fig. P13.33.

13.42 Find the Fourier transform of the following signals with A = 2, ω0 = 5 rad/s, α = 0.5 s−1 , and φ0 = π/5. (a) f (t) = A cos(ω0 t − φ0 ), − ∞ ≤ t ≤ ∞ (b) g(t) = e−αt cos(ω0 t) u(t)

(a) Derive the expression for υout (t) using Fourier analysis.

13.43 Find the Fourier transform of the following signals with A = 3, B = 2, ω1 = 4 rad/s, and ω2 = 2 rad/s. (a) f (t) = [A + B sin(ω1 t)] sin(ω2 t) (b) g(t) = A|t|, |t| < (2π/ω1 ) 13.44 Find the Fourier transform of the following signals with α = 0.5 s−1 , ω1 = 4 rad/s, and ω2 = 2 rad/s. (a) f (t) = e−αt sin(ω1 t) cos(ω2 t) u(t) (b) g(t) = te−αt , 0 ≤ t ≤ 10α 13.45

Using the definition of Fourier transform, prove that F [t f (t)] = j

13.46

d F (ω). dω

Let the Fourier transform of f (t) be F (ω) =

A . (B + j ω)

Determine the transforms of the following signals (using A = 5 and B = 2).

(b) Plot υout (t) using A = 5, T = 3 ms, R2 = 2 k, and C = 0.33 μF.

R1 = 500 ,

(c) Repeat (b) with C = 0.33 mF and comment on the results. 13.50 The circuit in Fig. P13.19 is excited by the source waveform shown in Fig. P13.34. (a) Derive the expression for υout (t) using Fourier analysis. (b) Plot υout (t) using A = 5, T = 3 s, R2 = 2 k, and C = 0.33 mF.

R1 = 500 ,

Section 13-10: Multisim 13.51 Design a Sigma-Delta converter that converts a sinusoidal voltage input with a magnitude always ≤ |1 V| and generates a digital signal with 0–5 V range. No voltage into any op amp can exceed ±20 V. 13.52 Design a Sigma-Delta converter that converts a sinusoidal current input with a magnitude always ≤ |1 mA| and generates a digital signal with 0–5 V range. (Hint: The easiest way to do this is to add an additional op-amp buffer ahead of the subtractor input to convert the current signal into a voltage signal.) No voltage into any op amp can exceed ±20 V.

“book” — 2015/5/4 — 7:27 — page 725 — #52

PROBLEMS

725

Potpourri Questions 13.53

How is data rate related to bandwidth and SNR?

(c) Measure the maximum value of υo (t) from the plot, and also measure the first time at which the maximum value occurs after t = 0.

13.54 In brain-machine interfaces, what are the electrodes connected to?

υs(t)

Integrative Problems: Analytical / Multisim / myDAQ

A

To master the material in this chapter, solve the following problems using three complementary approaches: (a) analytically, (b) with Multisim, and (c) by constructing the circuit and using the myDAQ interface unit to measure quantities of interest via your computer. [myDAQ tutorials and videos are available .] on †m13.1 Fourier Series Representation: Consider the periodic voltage waveform υ(t) shown in Fig. m13.1. (a) Determine if the waveform has dc, even, or odd symmetry. (b) Obtain its cosine/sine Fourier series representation. (c) Convert the representation to amplitude/phase format and plot the amplitude line spectrum for n = 0 using A = 10 V and T = 4 ms.

−T

−T/2

t

2T

(a) Waveform R υs

R

_ +

C

υo

R

υ(t)

..

T

0

..

A

T/4

−T/4

T/2

(b) Circuit Figure m13.2 Voltage waveform for Problem m13.2.

t

Figure m13.1 Voltage waveform for Problem m13.1.

m13.3 Fourier Transform: For the waveform shown in Fig. m13.3: (a) Determine its Fourier transform.

m13.2 Circuit Applications: The sawtooth voltage waveform υs (t) shown in Fig. m13.2(a) with A = 5 V and T = 2 ms serves as the input to the circuit of Fig. m13.2(b). (a) Determine the Fourier series representation of υo (t). (b) Plot υo (t) and υs (t) with MathScript or MATLAB as follows: (1) Time 0 ≤ t ≤ 5 ms,

(b) Plot the amplitude |F(ω)| with MathScript or MATLAB as follows: (1) Frequency 0 ≤ f ≤ 4000 Hz (remember to convert angular frequency ω to oscillation frequency f ), (2) A = 10, and (3) τ = 1, 2, and 4 ms (create three distinct plots)

(2) Sum of nmax = 100 terms, and (3) Circuit components R = 5.6 k and C = 0.1 μF. Use sufficient time resolution to display Gibbs phenomenon ringing. † Complete solution available on

.

(c) Determine the frequency at which the first null occurs in each of the three plots. (d) Discuss the relationship between the rectangular pulse width and the width of the main lobe of the amplitude spectrum.

“book” — 2015/5/4 — 7:27 — page 726 — #53

726

CHAPTER 13

FOURIER ANALYSIS TECHNIQUE

R

f(t)

+

A

+

υs(t)

_

C

υo(t)

4T

5T

_

(a) Circuit −τ/2

0

τ/2

t

Figure m13.3 Rectangular pulse waveform for Problem m13.3.

m13.4 Circuit Analysis with Fourier Transform: The circuit of Fig. m13.4(a) is excited by the double-pulse waveform shown in Fig. m13.4(b). (a) Derive the expression for υo (t) using Fourier analysis. (b) Plot υs (t) and υo (t) on the same graph over the time span 0 ≤ t ≤ 5 ms with MathScript or MATLAB for the following values: A = 5 V, T = 1 ms, R = 5.6 k, and C = 0.1 μF. (c) Determine the value of υo (t) at time t = 2 ms and t = 3 ms.

υs(t) A

T

2T

3T

t (ms)

(b) Waveform Figure m13.4 Double-plus waveform and circuit for Problem m13.4.

“book” — 2015/5/4 — 7:32 — page 727 — #1

A

APPENDIX

Symbols, Quantities, and Units

Symbol

Quantity

SI Unit

Abbreviation

A A B C d E F F f G G G g H I, i k L

Cross-sectional area Op-amp gain Bandwidth Capacitance Distance or spacing Electric field Force Fourier transform Frequency Conductance Closed-loop gain Power gain MOSFET gain constant Transfer function Current Spring constant Laplace transform

meter2 dimensionless radians/second farad meter volt/meter newton (variable) hertz siemen dimensionless dimensionless amperes/volt (variable) ampere newtons/meter (variable)

m2 — rad/s F m V/m N (variable) Hz S — — A/V (variable) A N/m (variable)

“book” — 2015/5/4 — 7:32 — page 728 — #2

728

APPENDIX A

SYMBOLS, QUANTITIES, AND UNITS

Symbol

Quantity

SI Unit

Abbreviation

L L,  N P, p P pf Q, q Q Q R S S s T,t u V, υ W W, w X Y Z α α β β ε  λ λ μ ρ σ τ φ χe ξ ω

Inductance Length Number of turns Power Mechanical stress Power factor Charge Reactive power Quality factor Resistance Cross-sectional area Complex power Complex frequency Time Velocity Voltage (potential difference) Width Energy Reactance part of impedance Admittance Impedance Piezoresistive coefficient Damping coefficient Common-emitter current gain Air resistance constant Permittivity Magnetic flux linkage Time shift Wavelength Magnetic permeability Resistivity Conductivity Time constant or duration Phase Electrical susceptibility Damping factor Angular frequency

henry meter dimensionless watt newtons/meter2 dimensionless coulomb volt·ampere reactive dimensionless ohm meter2 volt·ampere radians/second second meters/second volt meter joule ohm siemen ohm meters2 /newton nepers/second dimensionless newtons·second/meter farads/meter weber second meters henrys/meter ohms/meter siemens/meter second radians dimensionless dimensionless radians/second

H m — W N/m2 — C VAR —  m2 VA rad/s s m/s V m J  S  m2 /N Np/s — N·s/m F/m Wb s m H/m /m S/m s rad — — rad/s

“book” — 2015/5/4 — 7:32 — page 729 — #1

B

APPENDIX

Solving Simultaneous Equations Electric circuit analysis methods help us solve for the unknown voltages and currents in the circuits: • Kirchhoff’s circuit law solves for unknown branch currents. • Node-voltage method solves for unknown node voltages. • Mesh-current method solves for unknown mesh currents. In order to solve for n of these unknown currents or voltages, we need n independent equations that relate known values (such as the resistors, voltage and current sources, and other elements in the circuit) to the unknown voltages or currents. Once we have these n equations, we can write them in standard matrix form and solve them using either Cramer’s rule (by hand) or matrix inversion (using MATLAB, MathScript RF Module, similar software solvers, or your engineering calculator). This Appendix provides a brief overview of these approaches.

B-1 Review of Cramer’s Rule Let us assume that the application of Kirchhoff’s current and voltage laws to a certain circuit led to the following set of equations:

to apply the method of elimination of variables. If we solve for i1 in Eq. (B.1c), for example, and then use the expression i1 = (i2 + 5) to replace i1 in Eqs. (B.1a and b), we end up with two new equations containing only two variables, i2 and i3 . Repeat of the substitution procedure leads to a single equation in only one unknown, which can be solved directly. Once that unknown has been determined, it is a straightforward process to solve for the values of the other two variables. Such a solution method might prove effective for solving a simple set of three simultaneous equations, but what if the circuit we wish to analyze happens to contain a large number of variables? In that case, the more expeditious approach is to take advantage of Cramer’s rule, whose implementation procedure is both systematic and straightforward. Our review of Cramer’s rule will initally use the set of three simultaneous equations given by Eq. (B.1) to demonstrate the mechanics of the solution procedure for a system of order 3. Afterwards, we will treat the general case of a system of order n (consisting of n independent equations in n unknowns).

B-1.1

System of Order 3

Step 1: Cast Equations in Standard Form

2(i1 + i2 ) − 10 + (3i2 − i1 − 4i3 ) = 0

(B.1a)

−3(i1 + i2 ) + 2(i1 + 3i3 ) = 0

(B.1b)

Before we can apply Cramer’s rule, we need to regularize the simultaneous equations into a standard system of the following form:

i1 − 5 − i 2 = 0

(B.1c)

a11 i1 + a12 i2 + a13 i3 = b1 ,

(B.2a)

Our task is to solve the three independent, simultaneous, linear equations to determine the values of the three unknowns, i1 to i3 . [Recall that independence means that none of the three equations can be generated through a linear combination of the other two.] One way to accomplish the specified task is

a21 i1 + a22 i2 + a23 i3 = b2 ,

(B.2b)

a31 i1 + a32 i2 + a33 i3 = b3 ,

(B.2c)

where the a’s are the coefficients of the variables, i1 to i3 , and the b’s are the unaffiliated constants. By expanding the bracketed

“book” — 2015/5/4 — 7:32 — page 730 — #2

730

APPENDIX B

quantities in Eq. (B.1) and collecting terms, we can convert the equations into the standard form defined by Eq. (B.2). Such a process leads to:

Each element in the determinant has an address j k specified by its row number j and column number k. Thus, a12 is in the first row (j = 1) and second column (k = 2). For the system given by Eq. (B.4),    1 5 −4   6  .  =  −1 −3 (B.7)  1 −1 0

i1 + 5i2 − 4i3 = 10, −i1 − 3i2 + 6i3 = 0, i 1 − i2

= 5.

(B.3a) (B.3b) (B.3c)

This generates a matrix equation of the form AI = B, where I is the vector of n unknowns (currents i1 through i3 in this case): ⎡ ⎤⎡ ⎤ ⎡ ⎤ 1 5 −4 i1 10 ⎣−1 −3 6 ⎦ ⎣i2 ⎦ = ⎣ 0 ⎦ . (B.4) 1 −1 0 5 i3 Note that a11 = 1, a21 = −1, and a33 = 0. The regularized set of three linear, simultaneous equations given by Eq. (B.4) is a system of order 3. Step 2: General Solution According to Cramer’s rule, the solutions for i1 to i3 are given by 1 ,  2 i2 = ,  3 i3 = ,  i1 =

(B.5a) (B.5b) (B.5c)

where  is the value of the characteristic determinant of the system represented by Eq. (B.4), and 1 to 3 are the affiliated determinants for variables i1 to i3 . The procedure for evaluating these determinants is covered in Steps 3 and 4. Before we do so, however, we should note that in view of the fact that  appears in the denominator in Eq. (B.5), Cramer’s rule cannot provide solutions for the unknown variables when  = 0. This is not surprising, because for any system of n unknowns, the condition  = 0 occurs when one, or more, of the equations is not independent. This means that the system contains more unknowns than the available number of independent equations, in which case it has no unique solution. Step 3: Evaluating the Characteristic Determinant The characteristic determinant is composed of the a-coefficients of the 3 × 3 system of equations:    a11 a12 a13     =  a21 a22 a23  . (B.6)  a31 a32 a33 

SOLVING SIMULTANEOUS EQUATIONS

To evaluate , we expand it in terms of the elements of one of its rows. For simplicity, we will always perform the expansion using the top row. The expansion process converts  from a determinant of order 3 into the sum of 3 terms, each containing a determinant of order 2. Expanding Eq. (B.7) by its top row gives  = a11 C11 + a12 C12 + a13 C13 = C11 + 5C12 − 4C13 , (B.8) where C11 , C12 , and C13 are the cofactors of elements a11 , a12 , and a13 , respectively. The cofactor of any element aj k located at the intersection of row j and column k is related to the minor determinant of that element by Cj k = (−1)j +k Mj k ,

(B.9)

and the minor determinant Mj k is obtained by deleting from the parent determinant all elements contained in row j and column k. Hence, M11 is given by  after removal of the top row and the left column,     a a23   −3 6  M11 =  22 = . (B.10) a32 a33   −1 0  For a determinant of order 2, expansion by the top row gives M11 = a22 M22 − a23 M23 = a22 a33 − a23 a32 ,

(B.11)

which is equivalent to diagonal multiplication of the upperleft and lower-right corners to get a22 a33 , followed with multiplication of the other two corners to get a23 a32 , and then subtracting the latter term from the former. Substituting the values of the coefficients we have M11 = (−3) × 0 − 6 × (−1) = 6.

(B.12a)

Similarly, M12 is obtained by removing from  the elements in row 1 and the elements in column 2,   a a23  M12 =  21 = a21 a33 − a23 a31 a31 a33  = (−1) × 0 − 6 × 1 = −6.

(B.12b)

“book” — 2015/5/4 — 7:32 — page 731 — #3

B-2

MATRIX SOLUTION METHOD

731

Finally, M13 is obtained by removing from  row 1 and column 3,    −1 −3   = (−1) × (−1) − (−3) × 1 = 4. (B.12c) M13 =  1 −1 

the solution for any variable ik of the system is

Inserting the values of the three minor determinants in Eq. (B.8) gives

where  is the characteristic determinant and k is the affiliated determinant for variable ik . Analogous with the 3 × 3 system of Section 3-1,  is composed of the a-coefficients:

 = C11 + 5C12 − 4C13

ik =

  a11   a21  = .  ..   an1

= M11 − 5M12 − 4M13 = 6 − 5 × (−6) − 4 × 4 = 20.

(B.13)

Step 4: Evaluating the Affiliated Determinants The affiliated determinant 1 for variable i1 is obtained by replacing column 1 in the characteristic determinant  with a column comprised of the b’s in Eq. (B.2). That is,      b1 a12 a13   10 5 −4     6  . (B.14) 1 =  b2 a22 a23  =  0 −3  b3 a32 a33   5 −1 0 Evaluation of 1 follows the same rules of expansion discussed earlier in Step 3 in connection with the evaluation of . Hence        −3 6  0 6  0 −3        − 5 − 4 1 = 10  −1 0  5 0 5 −1  = 10 × 6 − 5 × (−30) − 4 × 15 = 150. Application of Eq. (B.5a) gives i1 =

1 150 = = 7.5.  20

Similarly, 2 is obtained from  upon replacing column 2 with the b-column, and 3 is obtained from  upon replacing column 3 with the b-column. The procedure leads to 2 = 50, 3 = 50, i2 = 2 / = 50/20 = 2.5, and i3 = 3 / = 2.5.

B-1.2

For a regularized system of linear simultaneous equations given by a11 i1 + a12 i2 + a13 i3 + · · · + a1n in = b1 ,

(B.15a)

a21 i1 + a22 i2 + a23 i3 + · · · + a2n in = b2 ,

(B.15b)

.. .

.. .

an2

a13 a23 .. .

an3

(B.16)

··· ··· ···

 a1n  a2n  ..  , .  ann 

(B.17)

and k is obtained from  by replacing column k with the b-column. For example, 2 is given by   a11   a21  2 =  .  ..   an1

b1 b2 .. .

bn

a13 a23 .. .

an3

··· ··· ···

 a1n  a2n  ..  . .  ann 

(B.18)

To determine the value of a determinant of order n, we can carry out a process of successive expansion, analogous with that outlined in Step 3 of Section 3-1. The first step in the expansion process converts  from a determinant of order n into a sum of n terms, each containing a determinant of order (n−1). Each of those new determinants can then be expanded into the sum of determinants of order (n − 2), and the process can be contined until it reaches a determinant of order 1, which consists of a single element.

B-2

Matrix Solution Method

The system of three simultaneous equations given by Eq. (B.2) can be cast in matrix form as

System of Order n

.. .

a12 a22 .. .

k , 

.. .

.. .

.. .

an1 i1 + an2 i2 + an3 i3 + · · · + ann in = bn ,



⎤⎡ ⎤ ⎡ ⎤ a11 a12 a13 i1 b1 ⎣a21 a22 a23 ⎦ ⎣i2 ⎦ = ⎣b2 ⎦ , a31 a32 a33 i3 b3

(B.19)

or in symbolic form as (B.15n)

AI = B,

(B.20)

“book” — 2015/5/4 — 7:32 — page 732 — #4

732

APPENDIX B

where

Each cofactor is a 2 × 2 determinant. Application of the definition given by Eq. (B.9) leads to



⎤ a11 a12 a13 A = ⎣a21 a22 a23 ⎦ , a31 a32 a33 ⎡ ⎤ i1 I = ⎣ i2 ⎦ , i3 ⎡ ⎤ b1 B = ⎣b2 ⎦ . b3

(B.22)

where A−1 is the inverse of matrix A. The inverse of a square matrix is given by A−1 =

adj A , 

⎤ 6 4 18 adj A = ⎣6 4 −2⎦ . 4 6 2

(B.23)

(B.27)

Upon incorporating Eqs. (B.22) and (B.23) and using the value of  obtained in Eq. (B.13), we have ⎡ ⎤⎡ ⎤ ⎡ ⎤ 6 4 18 10 i1 1 ⎣6 4 −2⎦ ⎣ 0 ⎦ . I = ⎣i2 ⎦ = 20 4 6 2 5 i3

(B.21c)

Matrix A is always a square matrix (same number of rows and columns), so long as the system of simultaneous equations contains the same number of independent equations as the number of unknowns. The solution for the unknown vector I is given by I = A−1 B,



(B.21a)

(B.21b)

SOLVING SIMULTANEOUS EQUATIONS

(B.28)

Standard matrix multiplication leads to ⎡ ⎤ 10   1 1 6 4 18 ⎣ 0 ⎦ = i1 = (6×10+4×0+18×5) = 7.5. 20 20 5 (B.29) Similarly, multiplication using the second and third rows of adj A leads to i2 = i3 = 2.5.

B-3

MATLAB or MathScript Solution

where adj A is the adjoint of A and  is the determinant of A. The adjoint of A is obtained from A by replacing each element aj k with its cofactor Cj k , and then transposing the resultant matrix, wherein the rows and columns are interchanged. Thus,

In MATLAB or MathScript software, matrices A and B of Eq. (B.25) are entered as:

adj A = [Cj k ]T .

The solution of AI = B is obtained by entering the statement

(B.24)

To illustrate the matrix solution method, let us return to the three simultaneous equations given by Eq. (B.3). Matrices A and B are given by ⎡

⎤ 1 5 −4 A = ⎣−1 −3 6 ⎦ , 1 −1 0 ⎡ ⎤ 10 B = ⎣ 0 ⎦. 5

B = [10; 0 ; 5];

I = A \ B; The MATLAB or MathScript response would be I=

(B.25a)

7.5000 2.5000

(B.25b)

2.5000. More information on using MATLAB and MathScript is available in Appendix E.

According to Eq. (B.24), adj A is given by ⎤T ⎤ ⎡ C11 C21 C31 C11 C12 C13 adj A = ⎣C21 C22 C23 ⎦ = ⎣C12 C22 C32 ⎦ . C31 C32 C33 C13 C23 C33

  A = 1 5 −4; −1 −3 6; 1 −1 0 ;



(B.26)

“book” — 2015/5/4 — 7:33 — page 733 — #1

C

APPENDIX

Overview of Multisim by Joe Steinmeier For information on the acquisition of the Multisim software, see the book's website (http://CAD.eecs.umich.edu). Included on the book’s website is a brief tutorial for getting started with Multisim. Multisim is a useful program for simulating and analyzing circuits. While the textbook introduces many of Multisim’s fundamental concepts, in the interest of space, many others are left out. The tutorial strives to review as well as continue the textbook’s coverage of Multisim. Importantly, the demos help clear up common stumbling blocks and the sometimes strange idiosyncrasies of the Multisim software. The tutorial consists of 43 basic “Demos,” which are mixtures of problems, solutions, investigations, and experiments in Multisim. The Demos are distributed among Chapters 2–9 and also Chapters 12 and 13. Each demo attempts to focus on introducing one main concept of Multisim, although it is unavoidable that other concepts are introduced throughout. An index is included to allow for the quick referencing of the tutorial. The demos are intended to help you become proficient in Multisim via simple but powerful examples. As you study each chapter in class, it is a good idea to at least skim over them and do some of the sample problems on the book website. Multisim can be an invaluable tool when trying to understand how any circuit works! The demos, grouped by chapter, cover the following material:

Chapter 2 2.1 Introduction to Multisim/The Three-way Switch Reviews layout basics with a simple, yet elegant, circuit involving three-way switches and light bulbs.

2.2 Resistor Network Analysis Introduces the Multimeter tool and resistor circuits with many resistors. 2.3 Thermal Sensing Wheatstone Bridge Demonstrates a variable resistor sensor in a Wheatstone bridge circuit. 2.4 The Wattmeter in Multisim Introduces power measurement in resistive circuits using the Wattmeter tool. 2.5 A Study of Dependent Sources Discusses the various dependent sources in Multisim, their limitations and uses. 2.6 An Introduction to ABM Sources Discusses the Analog Behavioral Modeling (ABM) sources in Multisim, which allow for the creation of formula-based dependent sources. Chapter 3 3.1 DC Circuit Analysis I Discusses how to use the Measurement Probes and the Interactive Simulation mode to solve for the voltages and currents in simple circuits. 3.2 DC Circuit Analysis II Discusses how to use the DC Operating Point Analysis tool to solve for the voltages and currents in simple circuits.

“book” — 2015/5/4 — 7:33 — page 734 — #2

734

APPENDIX C

3.3 Multisim and Th´evenin and Norton Circuits Discusses a general technique, using the Measurement Probe, for determining Th´evenin/Norton equivalents of any circuit. 3.4 Maximum Power Transfer Uses the Wattmeter and the Interactive Simulation mode to examine power transfer in resistive circuits.

3.5 Plotting Power Transfer in Multisim Demonstrates how to make an ABM source behave like a time-varying resistor and uses this component in a Parameter Sweep analysis to plot power transfer as a function of a varying resistance. This demo presents a very useful technique for plotting how a DC output changes as a function of a changing device parameter. Chapter 4 4.1 Operational Amplifiers in Multisim Introduces the DC Operating Point Analysis tool and uses it to analyze an inverting op-amp circuit. Also provides very nice tips on making attractive, easy-to-read plots in Multisim. 4.2 Introducing the Function Generator and the Oscilloscope Introduces the Function Generator and Oscilloscope instruments and uses them to measure the voltage gain of an op-amp circuit. Also discusses the Interactive Simulation tool in more detail. 4.3 Introduction to Signal Sources and the Transient Analysis Introduces time-varying sources and the Transient Analysis tool in the context of a simple op-amp circuit. 4.4 Using an Operational Amplifier in a Simple Audio Mixer Combines the lessons of Demos 4.1, 4.2, and 4.3 to build a three-channel op-amp audio mixer and analyze its timedependent behavior. Chapter 5 5.1 Introduction to Transient Circuits Discusses how to build interactive switch-based transients in circuits with the Interactive Simulation tool.

OVERVIEW OF MULTISIM

5.2 Transient Analysis and First-Order Circuits Discusses how to build voltage-controlled switches to generate transients with the Transient Analysis tool. 5.3 Inductors in Multisim Extends Demos 5.1 and 5.2 to switch-driven transients in circuits with inductors.

5.4 Time Constants in RC Circuits Uses the Parameter Sweep tool to plot the response of an RC circuit as you vary the value of R. Chapter 6 6.1 Parallel RLC Circuit Analysis Applies the Oscilloscope and both types of switches discussed in Demos 5.1 and 5.2 to analyze the transient behavior of RLC circuits.

6.2 An Over-, Under-, and Critically Damped Circuit Applies ABM sources to modeling the three fundamental types of transient responses in an RLC circuit. Chapter 7 7.1 Measuring Impedance with the Network Analyzer A good introduction into the Network Analyzer tool. 7.2 Introduction to AC Analysis Discusses how to produce frequency response plots using the AC Analysis tool. The demo uses an RLC circuit as an example, but the technique is useful for any type of circuit. 7.3 AC Th´evenin Circuit Determination Uses the AC Analysis tool and the Network Analyzer instrument to determine the open circuit voltage and complex impedance of an RLC circuit at a specific frequency. Using this data, the demo shows how to calculate the Th´evenin equivalent circuit and demonstrates that the transient response of the original circuit and its Th´evenin equivalent are the same. 7.4 Making an Impedance Purely Real Uses the AC Analysis tool and the Network Analyzer instrument to adjust a circuit’s frequency response. 7.5 Modeling an AC-to-DC Power Supply Builds, tests, and analyzes a rectifier circuit, very similar to that in Section 7-9, using Multisim.

“book” — 2015/5/4 — 7:33 — page 735 — #3

APPENDIX C

OVERVIEW OF MULTISIM

7.6 Phase Shift Circuits in Multisim A good companion demo to Example 7-19 in Chapter 7. 7.7 The Logic Analyzer: An Introduction Describes the Logic Analyzer used in Example 7-19 of Section 7-10 in more detail.

735 9.3 Speaker Crossover Circuit (Plotting Multiple Filters at Once) This is a great companion demo to Technology Brief 18: Electrical Engineering and the Audiophile. It demonstrates how to design and test a basic audio crossover circuit with the AC Analysis tool. 9.4 AC Parameter Sweep in a Radio Tuner Circuit

Chapter 8 8.1 Introduction to RMS Values in Multisim Discusses how to use the Multimeter instrument to obtain root-mean-square (rms) values for voltage, current, and power in a circuit. 8.2 AC Power Using AC Analysis

Uses the Parameter Sweep tool and the AC Analysis tool to demonstrate how varying a capacitor’s value adjusts a filter’s response, thereby acting as a tuner (or “station selector”) for a radio receiver. 9.5 60-Hz Active Notch Filter Offers an analysis of a multi-stage op-amp bandstop (or notch) filter. A good companion demo to Sections 9-7 and 9-9.

8.3 Power Factor in Multisim These two demos (8.2 and 8.3) discuss the somewhat tricky business of plotting complex power and power factors as a function of frequency using the AC Analysis tool. Because of Multisim’s variable and equation nomenclature, a user can easily spend a long time trying to enter the right equations in the analysis tools. This demo clarifies the jargon! 8.4 Three-Phase in Multisim Demonstrates the three-phase source component in Multisim using the Measurement Probe and the Transient Analysis tool.

8.5 Maximizing Power Delivered to a Load in a Complex Circuit Uses the AC Analysis tool to calculate the power delivered to a load as a function of frequency. A good companion demo to Section 8-6.

Chapter 9 9.1 Introduction to Filters in Multisim Demonstrates how to use the Bode Plotter instrument and the AC Analysis tool to generate Bode plots of any circuit. 9.2 Modeling a (Very)-Low-Pass Filter in Multisim Models a real-life application of a low-pass filter. Frequency response and transient response are shown with the various Multisim analysis tools.

Chapter 12 12.1 Piecewise Linear Sources Provides more detail on the piecewise linear source. 12.2 Exponential Sources Provides more detail on the exponential source. Chapter 13 13.1 Introduction to the Spectrum Analyzer 13.2 Fourier Analysis in Multisim These two demos (13.1 and 13.2) describe how to measure a signal’s various frequency components using either the Spectrum Analyzer instrument and/or the Fourier Analysis tool.

13.3 Analysis of a Square Wave This demo uses the Spectrum Analyzer and Oscilloscope to demonstrate the construction of a square wave from superimposed sinusoidal components at different frequencies (i.e., the sum of the Fourier components).

“book” — 2015/5/4 — 7:33 — page 736 — #1

D

APPENDIX

Mathematical Formulas    x−y x+y cos 2 2     x+y x−y cos x − cos y = −2 sin sin 2 2 

cos x + cos y = 2 cos

D-1 Trigonometric Relations sin x = ± cos(x ∓ 90◦ ) cos x = ± sin(x ± 90◦ ) sin x = − sin(x ± 180◦ )

ej x = cos x + j sin x

cos x = − cos(x ± 180◦ )

sin x =

ej x − e−j x 2j

cos x =

ej x + e−j x 2

sin(−x) = − sin x cos(−x) = cos x sin2 x =

(Euler’s identity)

cos2 x + sin2 x = 1

1 (1 − cos 2x) 2

2π rad = 360◦

1 cos x = (1 + cos 2x) 2 2

1 rad = 57.30◦

sin(x ± y) = sin x cos y ± cos x sin y cos(x ± y) = cos x cos y ∓ sin x sin y 2 sin x sin y = cos(x − y) − cos(x + y)

D-2

constants)  sin ax dx = −

2 sin x cos y = sin(x + y) + sin(x − y) 2 cos x cos y = cos(x + y) + cos(x − y) sin 2x = 2 sin x cos x cos 2x = 1 − 2 sin2 x     x+y x−y sin x + sin y = 2 sin cos 2 2     x+y x−y sin x − sin y = 2 cos sin 2 2

Indefinite Integrals (a and b are

 cos ax dx =  

eax dx =

1 cos ax a

1 sin ax a

1 ax e a

ln x dx = x ln x − x 

xeax dx =

eax (ax − 1) a2

“book” — 2015/5/4 — 7:33 — page 737 — #2

D-3

DEFINITE INTEGRALS (M 

x 2 eax dx =



1 x sin ax − cos ax a2 a

x cos ax dx =

1 x cos ax + sin ax a2 a

x 2 sin ax dx =

2x a2x 2 − 2 sin ax − cos ax a2 a3

x 2 cos ax dx =

2x a2x 2 − 2 cos ax + sin ax 2 a a3

eax sin bx dx =

eax (a sin bx − b cos bx) a 2 + b2



 e

ax

eax cos bx dx = 2 (a cos bx + b sin bx) a + b2

e

ax

sin bx dx =







x 1 dx tan−1 = 2 +a a a    1 dx x 1 −1 x = + tan (x 2 + a 2 )2 2a 2 x 2 + a 2 a a  x 2 dx x = x − a tan−1 2 2 a +x a x2

x sin ax dx =



D-3

Definite Integrals (m and n are integers) 2π

eax 2 a + 4b2

0

0

π sin nx dx =

(a sin bx − 2b cos bx) sin bx +

2b2 a



0

0

π sin nx sin mx dx = 0,

2b2 (a cos bx + 2b sin bx) cos bx + a



cos nx cos mx dx = 0,



π sin nx cos nx dx = 0 a 2  = b2

0

π sin nx cos mx dx = 0

a = b

cos(a − b)x cos(a + b)x − , 2(a − b) 2(a + b)

sin2 ax dx = 

x sin 2ax − 2 4a

sin 2ax x cos2 ax dx = + 2 4a

n = m

0

2

2

⎧ ⎨0, ⎩

sin ax cos bx dx = −

n = m



cos ax cos bx dx =



π 2

0



sin(a − b)x sin(a + b)x + , 2(a − b) 2(a + b)

cos2 nx dx =

2



sin(a − b)x sin(a + b)x − , 2(a − b) 2(a + b)

cos nx dx = 0



eax cos2 bx dx = eax a 2 + 4b2

2π sin nx dx =

2

sin ax sin bx dx =



737 

eax 2 2 (a x − 2ax + 2) a3





AND N ARE INTEGERS)

a 2  = b2

2n , n2 − m 2

if m + n = even and m = n if m + n = odd and m = n

2π sin nx cos mx dx = 0 0

∞ 0

sin ax π dx = ax 2a

“book” — 2015/5/4 — 7:33 — page 738 — #1

E

APPENDIX

MATLAB and MathScript E-1

Background

MATLAB MATLAB is a computer program developed and sold by the Mathworks, Inc. “MATLAB” is an abbreviation for MATrix LABoratory. It was originally based on a set of numerical linear algebra programs, written in FORTRAN, called LINPACK. So MATLAB tends to formulate problems in terms of vectors and arrays of numbers, and often solves problems by formulating them as linear algebra problems.

MathScript: For information on the acquisition of MathScript, see the book's website (CAD.eecs.umich.edu). MathScript is a computer program developed and sold by National Instruments, as a module in LabView. The basic commands used by MATLAB also work in MathScript, but higher-level MATLAB commands may not work in MathScript. For the purposes of this Circuits book, all necessary commands work equally well in both MATLAB and MathScript. A student version of MathScript is available for free for students using this book (see http://c3.eecs.umich.edu/ for instructions). Access to MATLAB is not required to use this book. In this sequel, we use “M/M” to designate “MATLAB or MathScript.”

Some basic things to know about M/M: • Inserting a semicolon “;” at the end of a command suppresses the output; without it M/M will type the results of the computation. This is harmless, but it is irritating to have numbers flying by on your screen. • Inserting ellipses “. . . ” at the end of a command means it is continued on the next line. This is useful for long commands. • Inserting “%” at the beginning of a line makes the line a comment; it will not be executed. Comments are used to explain what the program is doing at that point. • clear eliminates all present variables. Programs should start with a clear. • whos shows all variables and their sizes. • M/M variables are case-sensitive: t and T are different variables. • save myfile X,Y saves the variables X and Y in the file myfile.mat for use in another session of M/M at another time. • load myfile loads all variables saved in myfile.mat, so they can now be used in the present session of M/M. • quit ends the present session of M/M.

Getting Started

.m Files

To use the student version of MathScript, download LabView and then select MathScript under “Tools.” We will use this font to represent typed commands and generated output. You can get help for any command, such as plot, by typing at the prompt help plot.

An M/M program is a list of commands executed in succession. Programs are called “m-files” since their extension is “.m,” or “scripts.”

“book” — 2015/5/4 — 7:33 — page 739 — #2

E-2

BASIC COMPUTATION

To write an .m file, at the upper left, click: File → New → m-file This opens a window with a text editor. Type in your commands and then type: File → Save as → myname.m Make sure you save it with an .m extension. Then you can run the file by typing its name at the prompt: >> myname. Make sure the file name is not the same as a MATLAB command! Using your own name is a good idea. You can access previously-typed commands using uparrow and downarrow on your keyboard. To download a file from a web site, right-click on it, select save target as, and use the menu to select the proper file type (specified by its file extension).

739 To enter a vector of consecutive or equally-spaced numbers, follow these examples: • [2:6] gives ans=2 3 4 5 6 • [3:2:9] gives ans=3 5 7 9 • [4:-1:1] gives ans=4 3 2 1 To enter an array or matrix of numbers, type, for example, B=[3 1 4;1 5 9;2 6 5]; This gives the array B and its transpose B’ ⎡

⎤ 314 B = ⎣ 1 5 9⎦ 265



⎤ 312 B  = ⎣1 5 6 ⎦ 495

Other basics of arrays:

E-2 Basic Computation E-2.1

Basic Arithmetic

• Addition: 3+2 gives ans=5 • Subtraction: 3-2 gives ans=1 • Multiplication: 2*3 gives ans=6 • Division: 6/2 gives ans=3

• ones(M,N) is an M × N array of “1” • zeros(M,N) is an M × N array of “0” • length(X) gives the length of vector X • size(X) gives the size of array X For B above, size(B) gives ans=3 3 • A(I,J) gives the (I,J)th element of A. For B above, B(2,3) gives ans=9

• Powers: 23 gives ans=8

E-2.3 Array Operations

• Others: sin,cos,tan,exp,log,log10

Arrays add and subtract point-by-point: X=[3 1 4];Y=[2 7 3];X+Y gives ans=5 8 7 But X*Y generates an error message. To compute various types of vector products:

• Square root: sqrt(49) gives ans=7 • Conjugate: conj(3+2j) gives ans=3-2i √ Both i or j represent −1; answers use i. pi represents π . e does not represent 2.71828.

E-2.2

Entering Vectors and Arrays

To enter row vector [1 2 3] and store it in A type at the prompt A=[1 2 3]; or A=[1,2,3]; To enter the same numbers as a column vector and store it in A, type at the prompt either A=[1;2;3]; or A=[1 2 3];A=A’; Note A=A’ replaces A with its transpose. “Transpose” means “convert rows to columns, and vice-versa.”

• To multiply element-by-element, use X.*Y This gives ans=6 7 12. To divide element-by-element, type X./Y • To find the inner product of X and Y (3)(2) + (1)(7) + (4)(3) = 25, use X*Y’ This gives ans=25 • To find the outer product of X and Y ⎡

⎤ (3)(2) (3)(7) (3)(3) ⎣(1)(2) (1)(7) (1)(3)⎦ (4)(2) (4)(7) (4)(3) This gives the above matrix.

use X’*Y

“book” — 2015/5/4 — 7:33 — page 740 — #3

740 A common problem is when you think you have a row vector when in fact you have a column vector. Check by using size(X); in the present example, the command gives ans=1,3 which tells you that X is a 1 × 3 (row) vector. • The following functions operate on each element of an array separately, giving another array: sin,cos,tan,exp,log,log10,sqrt cos([0:3]*pi) gives ans=1 -1 1 -1 • To compute n2 for n = 0, 1 . . . 5, use [0:5].2 which gives ans=0 1 4 9 16 25 • To compute 2n for n = 0, 1 . . . 5, use 2.[0:5] which gives ans=1 2 4 8 16 32

APPENDIX E

E-2.4

• reshape(A(:),2,3) Unstacks the column vector to a 2×3 array which, in this case, is the original array A. • X=[1 4 1 5 9 2 6 5];C=X(2:8)-X(1:7) Takes differences of successive values of X. In the present example, the command gives C=3 -3 4 4 -7 4 -1 • D=[1 2 3]; E=[4 5 6]; F=[D E] This concatenates the vectors D and E (i.e., it appends E after D to get vector F) In the present example, the command gives F=1 2 3 4 5 6 • I=find(A>2) stores in I locations (indices) elements of vector A that exceed 2. find([3 1 4 1 5]2)=0 sets to 0 all values of elements of vector A exceeding 2. A=[3 1 4 1 5]; A(A 0.

“book” — 2015/5/4 — 7:33 — page 742 — #5

742

APPENDIX E

If H(s) is proper but not strictly proper, the constant K is nonzero. It is computed using [R P K]=residue(B,A);[R P],K since K has size different from R and P. To find the partial fraction expansion of H(s) =

s2 + 8s + 9 , s2 + 3s + 2

use the command [R P K]=residue([1 8 9],[1 3 2]);[R P] K       3 −2 3 −2 gives , K=1 so R= , P= , from which we read 2 −1 2 −1 off 3 2 H(s) = 1 + + . s+2 s+1 Double poles are handled as follows: To find the partial fraction expansion of H(s) =

8s2 + 33s + 30 , s3 + 5s2 + 8s + 4

MATLAB AND MATHSCRIPT

use the command [R P]=residue([8 33 30],[1 5 8 4]); ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ 3 −2 3 −2 ⎣ ⎦ ⎣ ⎦ ⎣ 4 −2 4 [R P] gives , so R= , P= −2⎦. We then 5 −1 5 −1 read off 5 4 3 + + . H(s) = 2 s + 2 (s + 2) s+1 In practice, we are interested not in an analytic expression for h(t), but in computing h(t) sampled every Ts seconds. These samples can be computed directly from R and P, for 0 ≤ t ≤ T : t=[0:Ts:T];H=real(R.’*exp(P*t)); Since R and P are column vectors, and t is a row vector, H is the inner products of R with each column of the array exp(P*t). R.’ transposes R without also taking complex conjugates of its elements. real is necessary since roundoff error creates a tiny (incorrect) imaginary part in H.

“book” — 2015/5/4 — 7:33 — page 743 — #1

F

APPENDIX

myDAQ Quick Reference Guide by Nathan Sawicki∗ Note: The myDAQ board, which does not come with the book, can be purchased directly from National Instruments. NI myDAQ is a student instrumentation and data acquisition device that allows students to make electronic measurements from their PC (Fig. F-1). Usually the circuit is built on a breadboard, which allows for making electrical connections between components without permanently soldering the connections. myDAQ converts the PC into an electronic instrumentation laboratory containing a variety of voltage sources (DC, AC, pulse, etc.) with which to excite the circuit, and standard test equipment (multimeter, oscilloscope, etc.) for measuring the voltage and current responses of the circuit. The physical wire connections made between the circuit and the myDAQ are essentially the same as those one would make to a real source or test equipment, and the graphical displays on the PC are visually very much like those one would see on the real test equipment. This appendix is a quick reference guide for the myDAQ and how to use it. Additional material including video tutorials, are available on the book website http://c3.eecs.umich.edu/.

PC myDAQ

Figure F-1: myDAQ connected to a PC and an electric circuit. The PC is used to provide the voltage inputs to the circuit, as well as to measure its output voltages and currents.

F-1.2

F-1 Getting Started with the myDAQ Follow the directions that came with your myDAQ to (a) install the NI ELVISmx software (if you have a MAC instead of a PC, you will need to install a Windows emulator before installing the myDAQ software), (b) install the 20-pin connector in the myDAQ, and (c) connect the myDAQ to the PC.

F-1.1

USB Port

Connect the myDAQ to the PC using the USB cable that came with the MyDAQ. The light next to the USB port will indicate that the myDAQ is functional (Fig. F-2). ∗ This appendix was originally written by Mr. Sawicki, a fourth-year student at the University of Michigan, and later edited by the book authors.

Circuit

Source Ports

Install the 20-Screw Terminal Connector into the myDAQ (you may need to press hard to insert the connector). This will give you access to: (1) analog voltage sources (± 15 V (left) and +5 V (far right pin)), which are always “on,” (2) two grounds, AGND and DGND, which are connected internally in the myDAQ, (3) two analog output ports, A0 and A1, which can provide variable DC, AC, and other waveforms up to ±10 V peak, (4) a digital I/O port DIO (which we will not be using in this book), and (5) Audio In and Audio Out ports.

“book” — 2015/5/4 — 7:33 — page 744 — #2

744

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

USB port Light

Terminal connector Measurement ports

Volts VΩ

Figure F-2: myDAQ unit. The myDAQ does not provide a current source, so we will demonstrate how to build one later on in this appendix. To access any of these ports, use the screwdriver that came with the myDAQ to loosen the screw at the top of the terminal, install the (stripped) end of the wire, and retighten the screw.  The ±15 V and 5 V sources are always “on” when the myDAQ is operational. The other sources are “on” or “off,” depending on the choices selected in the control panel on the PC display. Be careful not to short (touch) these sources to each other, to ground, to parts of your circuit where you do not intend them to be, or to you. Be careful when measuring within your circuit that metal probe tips do not accidentally touch (short) multiple points in the circuit. Although the current produced by the myDAQ is low, do not touch these sources (you can touch the insulated parts of wires, etc., but avoid touching the stripped ends and other bare metal parts attached to these sources). It is always a good idea when building or modifying circuits to disconnect the source from the circuit while making changes. 

F-1.3

Measurement Ports

The three measurement ports (Fig. F-2) provide access to the digital multimeter (DMM: voltmeter, ohmmeter, ammeter, as well as diode test). Depending on the type of measurement you wish to make, connect the probes that came with the myDAQ to these ports. One probe (black) should always be connected to the COM port, which is the common/reference/ground node. Then, connect the other (red) probe to one of the HI ports:

Amps I

A

com Voltmeter

Ammeter

Figure F-3: DMM ports. Note that these three connections represent the same three connections on the DMM graphic shown throughout the book.

(a) The left port is for voltage, resistance, and diode measurements, and (b) the right port is used for current measurements. Figure F-3 shows the proper connection for voltage and resistance measurements. The connection for current uses the right red port instead of the left one.

 WARNING: The maximum allowable myDAQ voltage is 60 VDC/Vrms. DO NOT plug the myDAQ into circuits with hazardous voltages, such as wall outlets and car batteries; doing so could damage the myDAQ or cause injury. READ the NI myDAQ User Guide and Specifications prior to NI myDAQ use. Be careful when using the ammeter to connect it in SERIES with the current being measured, NOT in parallel, or you could blow out the fuse on the myDAQ. 

F-1.4

NI ELVISmx Instrument Launcher

Run the NI ELVISmx software. If the instrument launcher is not already open, go to StartAll ProgramsNational InstrumentsNI ELVISmx for NI ELVISMX & NI myDAQNI ELVISmx Instrument Launcher.

“book” — 2015/5/4 — 7:33 — page 745 — #3

F-2

MEASURING RESISTANCE

745

Figure F-4: Instrument launcher. The instruments used in this book include (Fig. F-4): DMM: (Digital Multimeter): Measures magnitudes of voltage, current, and resistance.

F-2

Measuring Resistance

Bode: Measures the frequency response of a circuit.

1. Software: Run the NI ELVISmx software. Select the digital multimeter (DMM) from the instrument launcher. Select the ohmmeter, and its setup as shown in Fig. F-5. Select either Auto Range (a good start) or a range slightly above the value you expect to measure. Follow the instructions in step 3 below to connect the resistor you want to test to myDAQ, then press RUN. It will take a few seconds from when you connect the resistor to when the correct value is shown on the PC display.

ARB (Arbitrary Waveform Generator): Source for arbitrary waveforms including variable DC voltages.

2. Connection: Connect probe cables to the left HI and COM ports on the myDAQ as shown.

Scope: Measures time-varying voltages. FGEN (Function Generator): Source for time varying (oscillating) voltages including sine waves, square waves, rectangular functions, and pulse trains.

Measuring a 1 kΩ resistor

Auto or Specify Range

(a) Select DMM on Instrument Launcher Figure F-5: Using DMM to measure resistance.

(b) PC display

“book” — 2015/5/4 — 7:33 — page 746 — #4

746

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

Figure F-6: Connecting probes to a resistor.

3. Measure: Touch the pointy ends of the probes to either end of a resistor (red/black polarity does not matter). It is safe to hold the resistor and probes in your fingers. The probes are also pointy enough to poke directly into the protoboard holes, plus they have small gouges to help steady them against wires, as shown in Fig. F-6. For easier connection, you can use additional clips to connect the probes to your wires. It is important to have good quality connections between the measurement probes and your circuit, which requires steady pressure or additional clips. The resistor in the figure (Brown-Black-Red) is a 1 k resistor. NOTE: If you are measuring a resistor in a circuit, your measurement will include the effect of other connected circuit elements. Remove the resistor from the circuit if you are trying to measure its resistance alone.

terminal, as shown in Fig. F-8(b). Figure 1-18(b) of the text shows the connections for measurement of differential voltages and node voltages. Measure AA Battery: Touch the red DMM probe to the positive battery terminal while touching the black DMM probe to the negative battery terminal. Press hard enough to make a good electrical connection. You should see the DMM reading change to a value close to the battery’s voltage rating. AAA, AA, C, D cells should all give a reading of around 1.5 V. Now see what happens if you switch the leads and put the red lead on the − terminal and the black lead on the + terminal.

F-3 Measuring Voltage Specification: The maximum voltage the myDAQ can be connected to is 60 V DC/Vrms. Do not connect the myDAQ to a voltage higher than this. Software: Run the NI ELVISMX software. Select the DMM option (Fig. F-7). Connection: Connect probe cables to the leftr HI and COM ports on the myDAQ as shown in Fig. F-6. When connecting the voltmeter to the circuit, the red lead should be connected to the (+) terminal of the voltage being measured, and the black lead should be connected to the (−)

Measure myDAQ Voltages: Now let’s measure the voltages that come out of the myDAQ. Press RUN on the DMM, so that it is measuring voltage. You should see 0.0 V on the PC display until you connect the probes across an electrical device or a circuit. Press the red myDAQ probe to the screw on the +15 V

“book” — 2015/5/4 — 7:33 — page 747 — #5

F-4

MEASURING CURRENT

747

(a) Select DMM on Instrument Launcher

(b) PC display

Figure F-7: Using DMM to measure voltage.

connection on the 20-pin terminal and the black probe to the AGND pin.You should see a voltage close to +15 V. Repeat for the −15 V and +5 V measurements. What happens if you switch the + (red) and − (black) DMM leads in your measurements? Press the Stop button when you have finished your readings. Evaluate: Was the voltage you measured from your battery lower than its voltage rating? If so, why? What happened when you switched the leads?

(a) Probe leads

F-4

Node Va Vs

+ −

R1 Node Vb R2

Measures Vab

Measures Va (relative to ground)

(b) Voltmeters connected to measure voltage difference Vab and node voltage Va (relative to ground) Figure F-8: Connecting probes to measure voltage.

Measuring Current

Specification: The maximum current the myDAQ can be connected to is 1A DC/Arms. Do not connect the myDAQ to a current higher than this. Software: Run the NI ELVISMX software. Select the DMM software. Select the DC ammeter ( ), and its setup as shown in Fig. F-9. Select either Auto Range (a good start) or a range slightly above the value you expect to measure (20 mA is a good range for this test). Press RUN. The DMM is now measuring current (which will be 0.0 until you connect the probes in series with a current you want to measure). Connection: Connect the Probes to the myDAQ as shown in Fig. F-10. When connecting the ammeter to the circuit, the red lead is where the current comes in (the “tail” of the current arrow), and the black lead is where the current leaves (the “tip/ head” of the arrow). Figure F-10(b) shows the connections. If

“book” — 2015/5/4 — 7:33 — page 748 — #6

748

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

(a) Select DMM (a) Probe connection Measuring current through a 1 kΩ resistor with a 15 V source

Volts

Voltmeter

Amps

V12

V

Auto or Specify Range

+

1

R

2

I

Ammeter

− (b) DMM connections for measuring current

Figure F-10: Probe connections for measuring current.

(b) PC display Figure F-9: Using DMM to measure current.

the display result is negative, it means the current is flowing in the opposite direction. Measuring I : Since the myDAQ does not have a built-in current source (current sources are not standard sources), we need to build a simple circuit and measure the current flowing through it. To that end, build the circuit in Fig. F-11(a) with V = +15 V (from the myDAQ) and R = 1 k: • Connect a red wire (with both ends stripped) to the myDAQ +15 V and a similar black wire to the myDAQ AGND (Fig. F-11(b)). Tighten the screws in the 20-terminal connector to hold them securely in place.  The myDAQ is on, and the voltages are live, so don’t let the ends of the red and black wires touch each other.  • Insert a 1 k resistor (a resistor with brown-black-red lines) in your protoboard as shown in Fig. F-11(b). In the

protoboard, Row 2 corresponds to node 1 in Fig. F-11(a). The other end of the resistor (row 8) is node 2. • Connect the red wire from the myDAQ +15 V output to another point on row 2 on the protoboard. This is also node 1 in Fig. F-11(a). • Connect the red probe from the myDAQ to the other end of the resistor, using an alligator clip to make a good connection. This is node 2. • Connect the black wire from AGND to another row in the protoboard. This will be node 3. Connect a small piece of wire onto which you can clip the black alligator lead in this row also. • Connect the black probe from the myDAQ ammeter (COM) to the black wire alligator clip connected toAGND (node 3). Connect the red probe from the myDAQ ammeter (right-HI) to the other end of the red alligator clip (node 2). • You have now completed the circuit, and (ideally) the display is I = 15 V/1 k = 15 mA.  NOTE: Unclip the leads when you finish your measurement, as prolonged current will heat up the resistor. 

“book” — 2015/5/4 — 7:33 — page 749 — #7

F-4

MEASURING CURRENT

749

Node 1 1 kΩ

+ 15 V _

Node 2 Ammeter (a) Circuit diagram

(a) DMM connections

(b) Circuit connections (b) Measuring voltage Figure F-11: Measuring current through a 1 k resistor. Figure F-12: Measuring current by applying Ohm’s law. Measuring current with ammeter: On the DMM panel on the PC, press Run. You should see the DMM reading change to a value close to the expected current. Repeat using 5 V from the myDAQ. Now see what happens if you switch the ammeter probes. Once you have taken your readings, hit the Stop button and unclip the leads. Current flows so long as the leads are connected, and prolonged current will heat up the resistor.  If you accidentally try to make a voltage or resistance reading with the ammeter probe connections instead of the voltmeter connections, you can blow out the fuse on the myDAQ. Plan ahead. If you are not planning to make additional current measurements, return the probes to their voltmeter connections.  Measuring current with voltmeter: A very common way to “measure” the current is to measure the voltage and use Ohm’s law to calculate the current. This is shown in Fig. F-12. • Return the DMM probes to the voltmeter position. • Build the circuit in Fig. F-11(a) with a +15 V myDAQ source and 1 k resistor as shown in Fig. F-12. As soon as

you connect the red and black wires and insert the resistor, current starts to flow, so don’t leave it this way too long or the resistor will heat up. { Connect one end of the resistor to +15 V in row 2, which corresponds to node 1 in Fig. F-12(a). { Connect the black wire from the myDAQ AGND to the other end of the resistor. (this will be node 2 in Fig. F-12(a)). • Measure the voltage from node 1 (red voltmeter probe) to node 2 (black voltmeter probe). • Remove the resistor and then measure the resistance between nodes 1 and 2. • Determine the current I = V /R. Once you have taken your readings, hit the Stop button and disconnect the voltage (+15 V) from the circuit. Evaluate: Were the currents you measured higher or lower than expected? If so, why? What happened when you switched the DMM leads? How similar were the current readings using the ammeter with those using the voltmeter/Ohm’s law?

“book” — 2015/5/4 — 7:33 — page 750 — #8

750

APPENDIX F

Debug: If current = 0 There are several reasons you might have read zero current: • Did you make solid connections between power/ground and the circuit? Are the connections in the 20-pin socket tight, and did they make good connection to the resistor. If the voltage measured across the resistor is 0, the answer is “probably not.” • Did you connect to the circuit properly? Look very carefully at the picture of the connections to the resistor. Note that the connections on the protoboard need to be in the same row as the resistor. Read about the protoboard in the next section. • Did you make solid connections to the measurement probes? Are they in the correct DMM locations, depending on if you are measuring volts, ohms, or amps? Did you make a solid connection to the circuit? • Is the myDAQ plugged in and running (probably, or you would have noticed this by now). • Is something broken . . . possibly the fuse? Debug: If you think you might have blown the fuse Exceeding the NI myDAQ ammeter maximum current rating of 1 amp will most likely blow the protection fuse, in which

MYDAQ QUICK REFERENCE GUIDE

case the ammeter will always read zero and appear as an open circuit. To check the fuse, follow the video tutorial: https://decibel.ni.com/content/docs/DOC-12879.

F-5

Building with the Protoboard / Breadboard

A breadboard (also known as a protoboard or plugboard) is a base on which circuits are hand-built or prototyped. Breadboards allow connections between components and wires without permanently soldering the connection. They come in different shapes and sizes, but their basic connections are all the same. See the picture in Fig. F-13, showing the front (plug side) of the breadboard and the back side, which shows the metal clips that make up the nodes on the breadboard. (The back side is normally protected by a paper or plastic backing. If you remove this backing, the clips fall out, so just leave it in place.) The metal clips pinch onto the wires stuck into the plug side to make electrical connections. Each horizontal clip on the breadboard creates a node, so each ROW of the board is one node, and any wires plugged in on that row are connected to the same node. The vertical clips on the sides of the board create nodes that extend the full length of each side of the board. They are marked with red and blue lines and are often called rails. These rails are commonly used for power and ground by plugging a voltage (such as from the myDAQ)

Figure F-13: Breadboard front (plug side) and back (metal clips create nodes).

“book” — 2015/5/4 — 7:33 — page 751 — #9

F-6

USING THE NI MYDAQ AS A CURRENT SOURCE

751

Battery

+

_

(a) Series circuit Figure F-14: 8-pin dual-in-line package (DIP) chip plugged in across the divider on the breadboard. Each of the 8 pins is in a separate row (node). A small dot on the upper left corner of the chip, or a small dibet in the top of the chip, indicates which side is up.

on the red rail and ground (such as AGND from the myDAQ) on the blue rail. This makes power and ground convenient to the many places it is needed throughout the breadboard. The plastic center divider is just the right width to allow a dual-in-line package (DIP) chip to be plugged in with its legs in the holes on either side of this divider. We use this for op-amp chips (see Fig. F-14).

Battery

+

_

(b) Parallel circuit Figure F-15: Two light bulbs connected (a) in series and (b) in parallel.

Build example: Series and parallel resistors Figure F-15 shows light bulbs connected in series and in parallel, and Fig. F-16 shows an example of two 1 k resistors connected in series (between rows 40 and 50) and in parallel (between rows 30 and 35). Note that for the series combination, one end of each of the two resistors is plugged into row 45. Evaluate: Measure the resistance of the series and parallel pair of resistors. Build other combinations of these four resistors as well. How close were the measurements to what you expected? Connect +5 V across the series and parallel resistors. For the parallel combination, plug +5 V into row 30, and AGND into row 35. For the series combination, plug +5 V into row 40 and AGND into row 50. Measure the voltage across each resistor. Calculate the current through each resistor using Ohm’s law. Verify that series resistors have the same current through them, and parallel resistors have the same voltage across them.

F-6 Using the NI myDAQ as a Current Source The NI myDAQ cannot act as a stand-alone current source. However, we can use circuit elements to convert an input voltage into a steady current, thereby acting like a current source. One

Figure F-16: Protoboard connections for two resistors in series and in parallel.

such circuit element is the LM317 regulator. This tutorial explains how to use the LM317 to generate a current source. Using the LM317 voltage regulator The LM317 has three wires: Vin, Vout , and VAdjust . There are several different packages for the LM317, three of which are shown in Fig. F-17.

“book” — 2015/5/4 — 7:33 — page 752 — #10

752

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

Vout

Radj

Vin Vadj

Vout

Iout

Vin

ADJ

Figure F-18: Circuit configuration for building a current

Vout

source using the LM317 regulator.

(a)

(b) TO-39 (H) Metal can package Input Adjustment Output

Case is output Bottom view NS Package Number H03A

• The current output comes from the VAdjust port, so connect one end of a wire there and the other end to the + node of your load. • Connect AGND to the ground node of your circuit. Example 1: Build the circuit in Fig. F-19(a). • Build the LM317 current source circuit shown in Fig. F-19(b). The circuit requires a 1.84 mA current source, so select Radj =

(c) Figure F-17: Three types of LM317 packages. Note that (b) has the flat side up. For the circular configuration in (c), the pin arrangement corresponds to a bottom view of the metal can.

1.25 = 679.3  (round to 680 ). 0.00184

Connect Iout to the 1 k load resistor. • Measure the voltage across the 1 k resistor. You should get V = (1.84A)(1 k) = 1.84 V.

Building a current source

Example 2: Easily adjustable current source.

Figure F-18 shows how to connect the LM317 to build a current source.

The current source of Example 1 is fine, except that it requires us to recalculate and purchase a new resistor every time we change the magnitude of the current source. If you want to use your myDAQ at home, this may not always be practical. Instead, replace the 680  resistor with a potentiometer, which you can adjust with a screw or dial. However, the potentiometer can be turned down to R = 0, which conceptually would make the current go to infinity. In reality, the myDAQ (and most sources) has an inherent current limit. The myDAQ can provide 32 mA from the +15 V source. Although the myDAQ will limit the current to about 32 mA, even if you try to draw more, it is bad practice and can potentially damage the source if you try

• Connect +15 V to Vin . • Calculate the adjustment resistor based on your desired current current output: Iout =

1.25 . Radj

• Connect Radj between Vout and VAdjust .

“book” — 2015/5/4 — 7:33 — page 753 — #11

F-7

CREATING WAVEFORMS WITH THE FUNCTION GENERATOR (FGEN)

a Iout

1.84 mA

R1

1.0 kΩ

0

(a) Desired circuit

753

use. The larger the potentiometer, the smaller the current. Also the larger the potentiometer, the more sensitive it is when you turn the dial (the harder it is to dial the current you want). So, we should use the smallest potentiometer that gives us the minimum current required. For Radj (max) = 1 k, the minimum current would be 1.25 Imin = = 1.25 mA, 1 k

which is small enough for any of the examples in this book. A 1 k potentiometer is therefore a good choice. Our final circuit should look like that shown in Fig. F-20. Storing the current source The current source configuration is small enough that it can be stored on a breadboard without being disassembled. It is recommended that you build the LM317 circuit once, neatly, perhaps trimming the leads of the LM317 to make it fit snugly against the board. This way, you can use it whenever the need arises to construct a circuit with a current source.

(b) Wiring configuration

F-7

Creating Waveforms with the Function Generator (FGEN)

A function generator is used for creating AC and periodic voltage waveforms. The NI ELVISMX function generator is capable of creating sinusoidal, ramp, and square wave sources. It can also perform a sweep over a range of frequencies. Specifications: The A0 output is limited to ±10 VDC or Vrms and a current maximum of 2 mA. Software: Run the NI ELVISmx software. Select the function generator FGEN from the instrument launcher. Press RUN to start a continual waveform or SWEEP to sweep sequentially through a range of frequencies. Select the settings for your waveform:

(c) Equivalent current source circuit Figure F-19: Current source circuit example.

Produces a sinusoidal waveform Produces a ramp waveform Produces a square wave

to drive it beyond its limit. Therefore, the minimum Radj you should use is: Radj (min) =

1.25 = 39 . 32 mA

To be safe, insert a 39  current limiting resistor in series with the potentiometer. Now let’s consider what size potentiometer to

Frequency Range: From 200 mHz to 20 kHz. Note: Period T (s) = 1/Frequency (Hz). Amplitude: Peak-to-peak (Vpp ) up to 10 V. Vrms (for a sine √ wave) = 2 Vpp . DC Offset: Adds VDC to the waveform, from −5 V to +5 V.

“book” — 2015/5/4 — 7:33 — page 754 — #12

754

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

LM371

+ _

Current limiting resistor (39 Ω) Iout

Potentiometer

Load

Voltmeter

Figure F-20: Current source with LM317 regulator and a potentiometer to adjust Iout .

(a) Select FGEN

(b) Waveform selection Figure F-21: Using the function generator.

(c) PC display

“book” — 2015/5/4 — 7:33 — page 755 — #13

F-8

MEASURE A TIME-VARYING VOLTAGE WITH THE OSCILLOSCOPE

755

Software: Run the NI ELVISmx software. Select the oscilloscope from the instrument launcher. If you are not sure what settings to use, the Autoscale button is a good start. Press RUN. Figure F-22: Sweep settings for function generator.

Channel Settings: Decide which Analog Input (AI = 0 and/or AI = 1) to use, and check the Enable box(es) for channels to be viewed. Scale Volts/Div (y axis): Select how large each division is appropriate along the y axis (volts/division). Vertical Position (DC offset): This feature introduces a vertical offset to create a better view of the waveform, if needed. Set to 0 initially. Timebase (x axis): Select how large each division is appropriate along the x axis (seconds/division).

Figure F-23: Analog output port for voltage output for function generator.

Duty Cycle (only available for square wave): % of time the waveform is HI. Sweep Settings: Sweep (step) through frequencies from Start Frequency through Stop Frequency in steps of size Step. The Step Interval (sometimes called dwell time) is the amount of time the function generator stays at each frequency (Fig. F-22). Signal Route: Select which Analog Output (AO = 0 or 1) channel the waveform will appear on. Connection: The voltage waveform you create is accessible from the Analog Output ports (AO = 0 or AO = 1) on the 20pin connector. Connect a red wire to the AO port (0 or 1), as shown in Fig. F-23. This will be the (+) side of your voltage source. Connect a black wire to the AGND port, and this will be the (−) side of your voltage source. Tighten the screws to hold the wires in place. Press RUN when you are ready to turn on the function generator.

F-8 Measure a Time-Varying Voltage with the Oscilloscope The oscilloscope is used to measure time-varying voltages (Fig. F-24). You can think of it as a voltmeter that measures signals as a function of time. The scope can measure either one or up to two individual channels simultaneously, which is often useful when comparing input and output voltages.

Trigger Type: Toggle for when the waveform appears on screen. (a) Immediate, displays the waveform instantly. (b) Edge, the waveform is displayed only if its magnitude is higher or lower than the Level (V) setting. This is used to stabilize your view of the waveform. Acquisition Mode: It is recommended that you Run Continuously. The Once setting will display one waveform without re-acquiring a new signal. Connection: Connect the analog inputs to the circuit to be measured. The scope connection works essentially the same as the voltmeter connection. There are two channels (AI = 0 and AI = 1). Each channel has a (+) connection that should be connected to the (+) node of the voltage to be measured, and a (−) connection to be connected to the (−) node of the voltage to be measured. Figure F-25 shows a red lead connected to 0+ and a black lead connected to 0−. These should then be connected across the voltage to be measured. Channels 0 and 1 can be used at the same time or separately. Click Run on the user interface and you are ready to view your oscillating voltage. Example: Measuring the AC voltage from the function generator Use the myDAQ Function Generator to create a 4 Vpp 100 Hz AC sine wave. Connect it to the Scope by matching the red and black wires in the two photos of Fig. F-26. You should obtain the waveform seen in the scope window in Fig. F-22(b). Evaluate: Experiment with all of the settings on the function generator and verify their magnitudes and time periods on the scope. Remember that the period = 1/frequency.

“book” — 2015/5/4 — 7:33 — page 756 — #14

756

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

(a) Select Scope

(b) PC display Figure F-24: Using the oscilloscope.

F-9 Creating a Variable DC Source with the Arbitrary Waveform Generator The arbitrary waveform generator is used to create variable DC voltage sources or user-defined arbitrary time domain waveforms. We will use the ARB generator to create a variable voltage source. The generator can output one or two distinct

waveforms.† This tutorial covers only the basics of the ARB generator. See the online information for more details.‡ Specifications: The A0 output is limited to ±10 VDC or Vrms and a current maximum of 2 mA. Software: Run the NI ELVISmx software. Select the arbitrary waveform generator (ARB) from the instrument launcher † If two waveforms are to run continuously, they must be the same length

in time and number of samples. ‡Video tutorial: https://decibel.ni.com/content/docs/DOC-12941

“book” — 2015/5/4 — 7:33 — page 757 — #15

F-10

MEASURING FREQUENCY RESPONSE WITH THE BODE ANALYZER

757

Click the Waveform Editor icon to begin creating your custom waveform (Fig. F-28 will appear). Create a waveform: • Set the Sample Rate. Remember this setting, or add it to the filename, because you will need to specify this as the Update Rate in the ARB user interface (Fig. F-27).

Figure F-25: Scope mode cable connections.

• Add as many segments as desired in your waveform. For the Variable DC Voltage example, only one segment is needed. • Set the time duration of each segment (10 ms for this example). • For each segment add a New Component. Specify the component from the Function Library, an expression, or sketch. For this example, use Function Library  DC Level (Offset = 1.0) to create a 1 V DC voltage.

(a) Function generator

(b) Scope

Figure F-26: Wire connections for displaying the waveform generator output shown in Fig. F-22(b).

(Fig. F-27(a)). The waveform window (Fig. F-27(b)) provides a view of available functions, just like the oscilloscope option. Select and enable the output channel(s). These can also be the A0 or audio output channels. Select a waveform for each channel you enable. (You will create a waveform file using the editor described below if you want to use a custom output.) Select the Update Rate on the waveform window to be the same as the Sample Rate on the editor window. Select Run Continuously or Run Once. Select the Gain. The system creates a variable DC voltage source by creating and saving a DC waveform with magnitude 1.0 and then adjusting the gain to create the desired DC voltage. Press RUN to start the generator. Use the Waveform Editor to create a custom waveform as described below. Start this waveform by selecting it from the Waveform Name folder, and then set the Gain. In this example, you create a 1 V DC voltage with a Sampling/Update Rate of 1 kHz named 1VDC 1kHz.wdt. To change its voltage level, set the Gain to whatever voltage is needed (2 V in this case). The voltage waveform (2 V DC in this example) will appear in the User Interface window, and you can then use it for other applications.

• You can add additional components to each segment. Specify if these are to be added (+), subtracted (−), multiplied (×), divided (/), or frequency modulated (FM). For the variable DC voltage, no additional components are needed. • Save your file as a *.wdt file. Choose a name that describes the waveform and include in it its sample rate. Use 1 VDC 1 kHz as the filename for this example. Connection: The waveform voltage you create is accessible from the Analog Output ports (AO = 0 or AO = 1) on the 20-pin connector (Fig. F-29). Connect a red wire to the AO port (0 or 1). This will be the (+) side of your voltage source. Connect a black wire to the AGND port; this will be the (−) side of your voltage source. Tighten the screws to hold the wires in place.

F-10

Measuring Frequency Response with the Bode Analyzer

A Bode analyzer is used to plot the frequency response of a circuit or system (known as a Bode plot).§ The frequency and phase of voltages in a circuit become very important when capacitors and inductors are used. When these or other frequency-dependent components are used, the circuit may act differently at different frequencies. A Bode plot yields two important pieces of information for any measurable voltage in a circuit: § http://www.ni.com/white-paper/11504/en/.

“book” — 2015/5/4 — 7:33 — page 758 — #16

758

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

(a) Select ARB

(b) PC display Figure F-27: Arbitrary Wave Generator (ARB).

(1) The magnitude of a specific voltage in the circuit over a specified range of frequencies. (2) The associated phase of said voltage over the same range of frequencies. Software: Run the NI ELVISmx software. Select the Bode Analyzer from the instrument launcher (Fig. F-30). The Bode analyzer effectively drives the circuit with a Vin frequency sweep from the function generator (AO = 0). It reads Vin in the stimulus channel (AO = 0 by default) and Vout in the response

channel (AI = 1). It then compares them and plots the Gain = Vout /Vin in either linear or log scale. This is the frequency response of the system. To demonstrate how this is done, let’s measure the frequency response of the voltage across the capacitor in the series RC circuit of Fig. F-31(a). The Bode Analyzer User Interface is shown in Fig. F-31(b). (1) Connect Vin and set up its frequency sweep. Vin will be generated at the AO = 0 output.

“book” — 2015/5/4 — 7:33 — page 759 — #17

F-10

MEASURING FREQUENCY RESPONSE WITH THE BODE ANALYZER

759

Figure F-28: Waveform Editor User Interface.

Peak Amplitude: The Bode analyzer sends out a signal at AO = 0. Peak amplitude selects the amplitude of this signal. Mapping: Choose whether the x axis (frequency) is graphed logarithmically or linearly. (2) Connect the stimulus (measurement) channel.

Figure F-29: Cable connections for ARB.

Stimulus Channel: This channel measures the stimulus. The myDAQ defaults to AI = 0 as the stimulus channel. Connect AI = 0+ in the same place as Vin + (node a) and AI = 0− at the ground of your circuit. (3) Connect the response (measurement) channel.

• Connect AO = 0 to the Vin + node in the circuit (node a). • Connect AGND to the ground node in the circuit. Start/Stop Frequency: Set the range of frequencies to be measured. Steps per decade: This specifies how many data points are measured per frequency decade (10–100 Hz is a decade, for example).

Response Channel: This channel measures the frequency response of Vout . The myDAQ defaults toAI = 1 as the response channel. Connect the AI = 1+ port to the (+) node of Vout (node b) and the AI = 1− port to the (−) node of Vout (ground, in this example). (4) Press Run.You should see plots of the magnitude and phase of the frequency response of Vout on the PC display.

“book” — 2015/5/4 — 7:33 — page 760 — #18

760

APPENDIX F

MYDAQ QUICK REFERENCE GUIDE

Figure F-30: Bode analyzer.

a

R1

b

1 kΩ υin

~ + −

1 μF

C1 0

+

υout

_

(a) RC circuit

(b) Bode analyzer display and user interface Figure F-31: (a) RC circuit and (b) its gain magnitude and phase plots.

“book” — 2015/5/4 — 7:33 — page 761 — #1

G

APPENDIX

Answers to Selected Problems Chapter 1 1.1 (b) 4 µA (d) 390 GV 1.3 ( c ) 4 pF

Chapter 2 2.1  ≈ 2 km 2.3 (b) R ≈ 1, 174  2.7 R = 6.41 ; i = 17.2 A

1.13 I = 18 A along −z

2.11 R = 2500 

1.14 ( c ) i(t) = 0.12e−0.4t (pA)

2.15 Ix = 2.43 A

1.15 (d) q(t) = 1.7t[1 − e−1.2t ] nC 1.16 (b) Q(1, 12) = 2.948 (C) 1.17 (b) i = 0 @ t = 3 s. 1.20 i = 4.8 A

2.17 I1 = 2 A, I2 = 1 A, I3 = 2 A, I4 = 1 A 2.19 Ix = 3.57 A; Iy = 2.86 A 2.23 P = 0.32 W 2.25 V1 = −6 V, V2 = 0, V3 = 6 V 2.29 I0 = 31 A

1.22 (b) i = 0

2.34 R = 3 k

1.23 (a) i = 0

2.36 Vx = 8 V

1.25 ( a ) V2 = 48 V

2.38 Req = 9 

1.26 ( a ) −4 V

2.41 ( a ) Req = 5.5 

1.29 I = 5 A, W = 432 kJ 1.31 ( a ) p(0) = 0.5 W; p(0.25 s) = 1.36 Pmax = 6 W 1.38 Vy = 1.2 V 1.40 Vz = 2.5 V

2.44 I = 2 A 2.49 I = 1.97 A 2.51 I = 3.8 mA 2.55 P = 40 W 2.59 P = 4 W

“book” — 2015/5/4 — 7:33 — page 762 — #2

762

APPENDIX G ANSWERS TO SELECTED PROBLEMS

2.61 Req = 10 

3.68 VTh = −7.6 V; RTh = 1.6 

2.63 ( a ) R3 = 1.5 

3.72 IN = 7.71 A, RN = 4.54 

2.66 H = 0.6 mm

3.74 IN = 0.217 A; RTh = 9.2 

2.67 I1 = 0, I2 = 0.1 A

3.77 VTh = 1 V, RTh = 2.4 

Chapter 3 3.1 I = 3 A 3.3 P = −120 W 3.5 VR = 7 V 3.7 Ix = −0.1 A

3.80 VTh = 3 V, RTh = 1.5  3.83 Pmax = 2.09 mW 3.85 Pmax = 10 mW 3.89 I0 ≈ IREF 3.93 Vout ≈ (RL /RE )Vin

3.11 P = −24 W

Chapter 4

3.13 Ix = 0.77 A

4.1 υo = −10 V

3.15 Ix = 8 A

4.3 υo = −10 V

3.17 Vx = 1.41 V

4.6 (d) G = −100

3.20 Ix = 0.151 A

4.9 Rf = 16 k

3.23 Vx = −2.85 V

4.13 G = RL (R1 + R2 )/[R1 (R3 + RL )]

3.26 V = 10 V

4.14 (b) Rf = 180 k

3.28 V = 5 V

4.16 RL = 4 k

3.30 Ix = −0.1 A

4.19 G = 0.33; −21 V ≤ υs ≤ 21 V

3.32 Vx = 4 V

4.21 −2 V ≤ υs ≤ 2 V

3.34 Vx = −3 V

4.24 υo = 6.5 V

3.36 Ix = 6.25 A

4.30 υo = (38 − 4υs ) V; 5.5 V ≤ υs ≤ 13.5 V

3.39 Ix = 8 A

4.35 υo = −3.23 V

3.42 Vx = 1.67 V

4.38 υo = 0.826 V, P = 0.34 mW

3.45 I0 = 0.6 A

4.40 ( a ) G1 = −6, G2 = −20

3.48 Ix = 2 A

4.46 υo = −[(R3 /R2 )(R1 + R2 )/(R1 + Rs )]υs

3.52 V1 = 25.5 V; V2 = 4.5 V

4.50 υo = 8.5υs

3.56 Vx = 1.5 V

4.52 υo = −5.19 V

3.58 I = 0.05 A

4.54 υs = −0.1 V

3.62 Vx = −2.094 V

4.56 G = 2 × 105 to 2 × 106

3.64 VTh = 4 V; RTh = 5.2 

4.60 υo = 2.5 − 104 υs

“book” — 2015/5/4 — 7:33 — page 763 — #3

763

Chapter 5

Chapter 6

5.3 υ(t) = 10u(t − 2 μs) − 10u(t − 7 μs) V

6.1 υC (0) = 12 V, iL (0) = 0, iC (0) = −3 A, υL (0) = 12 V, υC (∞) = 0, iL (∞) = 4 A

5.7 υo = 12 V; τ = 2 s.

6.3 ( a ) υC (0) = −12 V, iL (0) = 3 mA

5.9 υ(t) = 2 + 8e−0.5t V

6.7 i1 (0) = 1 A; i2 (0) = 2 A 6.10 iC (0) = −3 A, υC (0) = −24 V, iR (0) = −3 A, υR (0) = −24 V, iL (0) = 0, υL (0) = −24 V, υL (∞) = 0, iR (∞) = 0, υC (∞) = 0, iL (∞) = 0

5.15 υ1 = −12 V; υ2 = 6 V; υ3 = 2 V 5.17 Ceq = 4 μF

6.12 υC (t) = (12.64e−2.68t − 3.64e−9.32t )

5.19 Ceq = 2.95 F

6.14 iL (t) = −90e−6t A

5.21 8 μF: 60 V; 3 μF: 60 V; 6 μF: 20 V; 6 μF: 30 V; 12 μF: 10 V; 10 μF: 30 V 5.25 i(t) =

(0.25 − e−0.2t ) A

√ 6.16 υC (t) = V0 cos(ωd t); ωd = 1/ LC

6.18 iC (t) = −[(40/3)e−0.5t sin 0.375t] A 6.22 iC (t) = (12 sin t − 6 cos t)e−2t A

5.29 υC1 = 20 V; υC2 = 12 V; iL1 = 0; iL2 = 2 A

6.25 υC (t) = 0.4 + (2.44t − 0.2)e−5.81t mV

5.31 Leq = 5 mH

6.28 ( a ) υC (∞) = 10 V

5.33 ( c ) iC (∞) = 0; υC (∞) = 8 V

6.29 υC (t) = (12 + 3e−60t ) V 6.32 υC (t) = 4−e−6000t [10 cos(11431t)−5.25 sin(11431t)]V

5.36 υ(t) = 14e−11.67t V 5.39 υ(t) = [−18 + 24e−1.25t ] V

6.35 υC (t) =  −666.7t 4 20 V 3 + 3 cos(745.4t) + 5.96 sin(745.4t) e

5.43 υ(t) = 5.35e−1000(t−0.1)/45 V

6.38 iL (t) = 1.5 mA 6.40 iL (t) = [32 − (32 + 6400t)e−400t ] mA

5.45 υC (t) = 4e−100t/1.5 V

6.43 ( a ) iL (t) = −2 + [4.5 cos(526.8t) + 7.26 sin(526.8t)]e−850t A,

5.48 i(t) = 2e−100t/3 A 5.50 i1 (t) = 2.88e−10t mA; i2 (t) = 0.72e−20t mA

6.46 ( a ) iL (t) = 2 A

5.53 i(t) = 0.3(1 − e−8t ) A 5.56 i(t) =

[20 − 30e−500t ]

(b) No, the solution method is applicable to dc sources only.

mA

1 5.60 υout (t) = υout (0) + υi (t) + RC

t

5.63 iout (t) = 0.6e−2t (mA) 5.65 (b) υout (t) =

(b) wC (∞) = 0

−24(1 − e−2t ) V

υi dt 0

6.48  iC (t) = for 0 ≤ t ≤ 1 ms (−0.0045e−83t + 0.1045e−1917t ) A −3 (2.43 × 10 cos 400t − 0.017 sin 400t) A for t ≥ 1 ms 6.50 υ(t) = (24 − 14.4e−2t + 6.4e−3t ) V 6.52 ( c ) [3 cos 0.433t + 5.2 sin 0.433t]e−0.75t A 6.54 υC (t) = 2.4 − (0.4 cos 3.74t − 0.428 cos 3.74t)e−6t V

5.68 ( a ) υout1 (t) = −0.5t V

6.56 i2 (t) = (2.3e0.89t + 97.7e−5.61t ) μA

5.73 W = 0.2 μm

6.58 υout (t) = −8e−20t sin 74.83t V

“book” — 2015/5/4 — 7:33 — page 764 — #4

764

APPENDIX G ANSWERS TO SELECTED PROBLEMS

Chapter 7

7.58 iC (t) = 1.25 cos(400t − 6.35◦ ) A

7.1 A = 4 V, f = 4 kHz, t = 0.25 ms, φ = 45◦

7.60 V1 = (20.7 + j 16.1) V, V2 = −(20.7 + j 42.1) V

7.3 υ(t) = 12 cos(8π × 103 t + 60◦ ) V 7.6 t = 1/9 μs, waveform shifts backwards 7.9 υ(t) = 12 cos(100πt − 45◦ ) V ◦

7.10 ( c ) z3 = 7.21e−j 146.3 ( e ) z5 =

◦ 5e−j 53.13



7.62 IC = 1.93ej 4.9 A   3 + j1 Vs 7.65 Vout = 5 7.68 ix (t) = 24.72 cos(5 × 105 t − 74.06◦ ) A ◦

7.11 ( c ) z3 = −1.22(1 + j ) ( f ) z6 = 3 + j 4

7.71 VTh = 10e−j 30.5 V, ZTh = (2.9 − j 3)  7.74 ia (t) = 2.06 cos(35t + 152.75◦ ) A ◦

7.14 (b) z2 = 100e−j 73.74

7.77 Ia = 4.4e−j 21.45 A



7.80 υout (t) = V0 sin ωt

( f ) Im(z∗ ) = −6 7.15 ( c ) z1 z2∗ = 10e−j 105

7.83 υout (t) = 11.32 cos(377t + 152.05◦ ) V

7.16 (b) ez = −2.45 − j 2.24

Chapter 8



7.18 (b) B = 42.5e 7.20 (b) 62.2e

−j 161.99◦

−j 4.61◦

(b) Vrms = 2.31 V ◦

8.4 ( a ) Iav = 3 A

7.22 (b) V2 = 2ej 108 V 7.23 ( c ) i2 (t) = 2 cos(2π × 103 t + 150◦ ) A

7.27 (b) υ2 (t) = 5.49 cos(1000t − 18◦ ) V 7.29 iC (t) = 9.4 cos(2π × 104 t − 21.48◦ ) mA 7.32 υab (t) = 0.42 cos(300t

− 186.35◦ ) V

7.34 (b) Z2 = (98.5 + j 1524)  7.36 Z = (5.32 − j 1.69)  j 35.56◦

(b) Irms = 3.46 A 8.7 ( a ) Vav = 2 V

7.25 ( c ) ZC = −j 3.18 

7.38 IR = 2.528e

8.1 ( a ) Vav = 2 V

A

7.40 ( a ) Z = (5 + j 5)  (b) I = 3.54 A 7.44 L = 0 or 2.5 mH 7.48 VTh = −12 V; ZTh = 0

(b) Vrms = 3.79 V 8.10 ( a ) Vav = 2 V (b) Vrms = 2.38 V 8.12 ( a ) Vav = 0.432 V (b) Vrms = 0.5136 V 8.15 ( c ) Vav = 12 V, Vrms = 12.32 V ◦

8.17 ( c ) S = 330ej 15 VA, Pav = 318.76 W, Q = 85.41 VAR, pf = 0.97 (lagging) ◦

8.18 ( c ) S = 2.5e−j 75 VA, Pav = 0.65 W, Q = −2.415 VAR, pf = 0.26 (leading) ◦

8.21 S = 0.665e−j 79.35 VA, pf = 0.185 (leading)

7.50 ZL = (6 + j 2) 

8.23 Pav (200 ) = 5.52 W, Pav (100 ) = 1.38 W, Pav (source) = 6.9 W

7.51 ZL = (2 − j 1) k

8.26 Pav = 496.4 mW

7.56 f = 795.8 Hz

8.29 PL = 0.186 W

“book” — 2015/5/4 — 7:33 — page 765 — #5

765 8.31 Sload = (0.38 + j 0.26) VA 8.33 PL = 3.933 mW 8.36 PL = 0.713 W 8.39 Pav = 0.2 mW 8.43 Zeq = (29 + j 8.93) , inductive Z1 = inductive, Z2 = capacitive, Z3 = inductive 8.45 C = L/(R 2 + ω2 L2 ) 8.48 ( a ) Z1 = (10.5 − j 5.2) ; Z2 = (7.23 + j 5.05) 

9.35 ( a ) H(ω) =

1 2

rad/s 9.37 ( a ) H(ω) =



 j ω/ωc , with ωc = 1.25 × 104 1 + j ω/ωc

1 + j ωC(R1 + R2 ) 1 + j ωR1 C

1 1 + j 0.4ω/5000 + (j ω/5000)2  −j (ω/ωc1 ) 9.41 ( a ) H(ω) = , with (1 + j ω/ωc2 )(1 + j ω/ωc3 )

9.39 ( a ) H(ω) =

ωc1 = 104 rad/s, ωc2 = 200 rad/s, ωc3 = 2 × 106 rad/s

8.50 ( a ) pf = 0.673 lagging

9.46 H(ω) = −10

8.52 ZL = (0.6 − j 0.2) , Pmax = 6.78 W

9.50 78 to 98 MHz, or 98 to 118 MHz

8.55 ZL = (1.33 − j 4) , Pmax = 4.18 W

Chapter 10

8.58 ZL = (5.1568 + j 0.72) , Pmax = 0.8 W

10.4 V1 = 100 −60◦ V (rms), negative phase sequence

8.60 ZL = (22.9 + j 39.8) , Pmax = 38.9 mW

10.7 υab (t) = 34.79 cos(377t + 113.06◦ ) V υbc (t) = 81.58 cos(377t + 47.78◦ ) V υca (t) = 101.2 cos(377t − 114.02◦ ) V

Chapter 9 9.1 ω0 = 104 rad/s 9.3 ω0 = 105 rad/s 9.5 ω0 = 5 × 104 rad/s 9.8 Km = 50; Kf = 2 × 104

10.11 iL1 (t) = 22.73 cos(2πf t + 16.9◦ ) A iL2 (t) = 27.71 cos(2πf t − 100.5◦ ) A iL3 (t) = 26.56 cos(2πf t + 129.0◦ ) A 10.14 iL1 (t) = 27.31 cos(2πf t + 7.62◦ ) A iL2 (t) = 32.71 cos(2πf t − 114.2◦ ) A iL3 (t) = 29.54 cos(2πf t + 117.5◦ ) A

9.11 (b) ω0 = 1 rad/s

10.17 IL1 = 10.02 −59.3◦ A (rms)

9.13 (b) −23 dB

10.20 In = 1.64 143.2◦ A (rms)

9.14 ( a ) 20.81 dB

10.23 ST = 29.68 −3.56◦ kVA

9.16 ( c ) 0.25 −2000ω 9.21 H(ω) = (1 + j ω)(100 + j ω) 9.23 H(ω) = −(10 + j ω)(100 + j ω)/50ω 9.26 ( a ) ω0 = 104 rad/s, Q = 40, B = 250 rad/s, ωc1 = 9875 rad/s, ωc2 = 10125 rad/s

10.26 ST = 29.68 −3.56◦ kVA, 58.2% 10.28 IL1 = 25.61 −28.37◦ A (rms) IL2 = 25.61 −148.37◦ A (rms) IL3 = 25.61 91.69◦ A (rms) 10.30 IL1 = 69.62 −19.1◦ A (rms) IL2 = 69.62 −139.1◦ A (rms) IL3 = 69.62 100.9◦ A (rms) pf = 0.9449

9.29 ω0 = 5 rad/s; Q = 20; B = 25 rad/s; ωc1 = 487.5 rad/s; ωc2 = 512.5 rad/s

10.32 C = 2.261 mF

j 10−3 ω 9.32 (b) H(ω) = 1 + j 2.5ω/1000 + (j ω/1000)2

10.36 P1 = P2 = 50.82 W

10.34 P1 = 3.35 kW, P2 = 6.194 kW

“book” — 2015/5/4 — 7:33 — page 766 — #6

766

APPENDIX G ANSWERS TO SELECTED PROBLEMS

Chapter 11

12.22 iL (t) = (0.012e−0.13t − 0.81e−3.31t ) u(t) mA

11.1 ( a ) i(t) = 0.293 cos(120π t − 125.8◦ ) A

12.25 iL (t) = 30e−2t sin t u(t)

(b) P = 8.57 W

12.28 iL (t) = [2 + 7e−0.67t cos(1.43t − 1012.5◦ )] u(t) A

11.4 Vout = 2.629 164.4◦ V 11.7 Ix = 0.8622 −51.9◦ A 11.9 Vout = 0.8

−73.3◦

V

12.31 iL (t) = [1.25 + 2.58e−1.75t cos(0.97t − 61◦ )] u(t) A 12.34 υC (t) = 10(1 − e−0.25t ) u(t) − 10[1 − e−0.25(t−2) ] u(t − 2) V

11.11 Vout = 24.83 −155.6◦ V

12.37 iL (t) = (21.52e−2t − 112.14e−t − 5.38e−5t ) u(t) A

11.13 ( c ) Leq = 2 H

12.44 ( a ) υout (t) = 2[1 − e−10 t (1 + 104 t)] u(t) V 4

11.15 Zin = 18.28 −38.9◦  11.17 Zin = 14.03

30.8◦



11.19 Leq = 5.1 H 11.21 Zin = (0.3 + j 9.9)  11.24 Vout = 26.63 33.7◦ V 11.26 Ix = 24.25 −129◦ mA 11.28 PL = 0.83 W

13.5 ( a ) even and dc symmetry    ∞

160 3nπ nπ (b) f (t) = − cos cos (nπ )2 4 4 n=1   nπ t · cos 4 13.5 ( a ) Odd and dc symmetry ∞

10 (b) f (t) = [2 − cos(nπ )] sin(nπ t) nπ n=1

11.31 PL = 2.676 W

13.12 f (t) =

11.33 n = 8/3

Chapter 12 12e−4s 12.4 ( c ) F3 (s) = s+3  48 12 12.5 ( a ) H1 (s) = e−4s + s + 3 (s + 3)2 60 ( c ) H3 (s) = (s + 2)4 12.7 ( c ) F3 (s) =

Chapter 13

10e−3s s+4

12.9 (b) f2 (t) = [2e−2t − 4e−3t cos t] u(t) 12.10 ( c ) f3 (t) = 2e−3t cos(2t + 45◦ ) u(t) 12.11 (b) f2 (t) = 4(1 + cos 3t) u(t) 12.12 iL (t) = 8(e−t − e−2t ) u(t) A 12.15 iL (t) = 1.14e−0.5(t−2) sin(0.87(t − 2)) u(t − 2) A 12.18 υ(t) = [1.5 − 1.56e−4t + 0.072e−12t ] u(t) V

1 2



1 2

cos(8π t)

  ∞ 40 1 nπ t 13.16 f2 (t) = − sin π n 4 n=1

13.26 ( a ) υout (t) =  2 ∞ nπ

4 100 × sin cos(nπ t + 90◦ ) V nπ 2 n=1

13.29 Pav = 54.5 mW; ac power fraction = 8.26% 5 [(1 + j 3ω)e−j 3ω − 1] 3ω2  3 2j 13.39 F(ω) = −2j − (sin ω − sin 2ω) ω ω 13.33 F(ω) =

13.46 (b) t f (t)

5 (2 + j ω)2

“book” — 2015/5/4 — 7:34 — page 767 — #1

Index

A

Amplifiers common collector, 178 common emitter, 178 operational, 183–247 Amplitude modulation (AM), 548 Amplitude/phase format, 681 Amplitude spectrum, 681 AM radio band, 548 Analog, 154–156 Analog circuits, 223–224 Analog signals, 154 Analog-to-digital converter (ADC), 154 Analysis techniques, 115–182 bipolar junction transistor (BJT), 158–161 by-inspection methods, 129–133 mesh-current method, 123–125, 128–129 nodal analysis with Multisim, 161–163 node-voltage method, 117–123 source superposition, 133–135, 140 Th´evenin and Norton circuits, 140–151 AND logic gate, 156, 539 Angular frequency, 386 Apparent power, 473 Arbitrary waveform generator, myDAQ, 756–757 Artificial eye retina, 363 Artificial sources, 117 ASCII, 154

Absolutely integrable, 704 ac analysis, 385–458 Accelerometer, 303 Acoustic touchscreens, 395 ac power, 459–499 average power, 463–464, 467 complex power, 467–472 maximum power transfer, 476, 481–482 power factor, 472–476 Active device, 2 Active digitizers, 395 Active filters, 536–538 Active matrix display, 191 Active noise control, 509 Active shutter 3-D, 638 ADC, 154 Adder, 200 Additivity property, 116 Admittance, 403 Air-core solenoid, 270 Alternating current (ac), 4, 23, 385, 386 American Wire Gauge, 52 Ammeter, 27–28 Ampere-hours, 29 767

“book” — 2015/5/11 — 17:43 — page 768 — #2

768 Audible spectrum, 544 Autotransformer, 618 Average ac power, 463–464, 467 Average power, 460, 693–694 Average real power, 582 Average value, 460 Average values of periodic waveforms, 460–462

B Backlight, 191 Balanced circuits, 83 Balanced condition, 85, 572–573 Balanced load, 573 Balanced networks, 573, 576, 579–582 Balanced source, 573 Balanced three-phase generators, 568–572 Bandpass filter, 501, 523–528 Bandreject filter, 502, 529–530 Bandwidth, 524, 688 Bandwidth data rate, 688–689 Base, 158 Biochemical pathways, 695 Bipolar junction transistor (BJT), 158–161 Bit, 154 Bit-patterned media (BPM), 294 Block, 140 Block diagram, 141 Bode diagram, 514 Bode plots, 512–522 Bouncy switch, 227 Brain stimulation, 363–364 Branch, 17 Bridge circuit, 157 Bridge rectifier, 433, 434 Buffer, 208 Bus, 306 Bus circuit, 568 Bus speed, 306 By-inspection methods, 129–133 mesh analysis by inspection, 132–133 nodal analysis by inspection, 130–132

C Cables, 545 Cantilever beam, 578 Capacitance, 258

INDEX Capacitive MEMS actuator, 338–339 Capacitive impedance, 403 Capacitive sensor/MEMS accelerometer, 336–337 Capacitive sensors applications, 301–303 Capacitive touch buttons, 301 Capacitive touchscreens, 393–394 Capacitors, 258–264, 268–269 electrical properties, 259–263 in the s-domain, 653 series and parallel, 263–264, 268–269 supercapacitors, 265–267 Carrier frequency, 547 Cascaded active filters, 538–43, 547 Cascaded systems, 149–151 Cathode ray tube (CRT), 190–191 Cell-phone circuit architecture, 2–4 Center-tapped pole transformer, 568 Characteristic equation, 342 Charge, 20–22 Charged capacitor, RC circuit, 276–279 Charge/discharge, 285 Charging-up mode, RLC circuit, 334–335 Chassis ground, 26 Circuit analysis by Laplace transform, 630–673 Multisim analysis, 662–664 partial fraction expansion, 644–647, 650–652 s-domain circuit analysis, 655–662 s-domain circuit element models, 652–654 Circuit analysis with Fourier transform, 711–713 Circuit breaker, 86 Circuit diagram, 15 Circuit elements, 35–40 Circuit equivalence, 67 Circuit gain, 186 Circuit representation, 15–20 Circuit response, 279 Circuit simulation software, 225–228 Multisim, 225–227 3-D modeling tools, 227–228 Circuit theory, 50 Clock speed, 277 Closed-loop gain, 186 CMOS (complementary MOS), 222 switching speed, 306–310 Coaxial capacitor, 259 Cochlear implant, 363 Cognitive prostheses, 363

“book” — 2015/5/4 — 7:34 — page 769 — #3

INDEX Cognitive radio, 713 Collector, 158 MEMS accelerometers, 337–338 Common collector amplifier, 178 Common-emitter amplifier, 159, 178 Compensated load, 474 Complementary MOS, 222 Complex conjugate, 390 Complex frequency, 634 Complex number, 389 Complex power, 460, 467–472, 582 Computer memory circuits, 203–205 Ferroelectric RAM (FeRAM), 205 Magnetoresistive RAM (MRAM), 205 Nano RAM (RAM), 205 random-access memories (RAMs), 204–205 read-only memories (ROMs), 203–204 Conductance, 69 Conductance matrix, 131 Conduction current, 21 Conductivity, 51 Conductors, 51 Conjugate matched, 481 Conjugate symmetry property, 710 Connected in parallel, 540 Connected in series, 539 Convergence condition, 634–635 Convergence of Fourier integral, 704 Corner frequency, 503 Cosine-referenced, 387 Cosmic rays, 465 Coupling coefficient, 611–612 Cramer’s rule, 729–731 Critically damped response, 346–348 Critical temperature TC, 57 Crossover circuits, 546 Crystal oscillators, 423–424 Cumulative charge, 23–24 Current, 22–24 Current coil, 591 Current-controlled voltage source, 39 Current divider, 74 Current division, 74–75 Current measurement, myDAQ, 747–750 Current mirror, 178 Cutoff frequency, 503

769

D Damped natural frequency, 348 Damping coefficient, 342 Damping factor, 517 dB scale, 512–515 dc gain, 503 dc magnetic field, 608 Deep brain stimulation (DBS), 363 Definite integrals, 737 Degree of selectivity, 524 Delay time, 663 -configuration, 586 Delta function, 631 -load configuration, 583 -source configuration, 570 Delta-Wye (-Y) transformation, 82–83 Demodulation, 548 Dependent source, 35, 38–40, 97–99 Dependent source circuit, 66–67, 120–121, 125 Dependent voltage source, 38 Deposition, 136 Design, 3 Destructive interference, 509 Dielectric, 51 Difference amplifier, 206–207 Differential measurement approach, 215 Digital and analog, 154–156 Digital inverter, 220–221, 231–234 Digital light processing (DLP), 194 Digital light projector (DLP), 339 Digital signal, 154 Digital-to-analog converter (DAC), 155, 216–219 Digital wattmeter, 591 Dimension, 9 Diode, 87–91 DIP configuration, 184 Direct current (dc), 4, 22 Dirichlet conditions, 677 Discharging mode, RLC circuit, 335, 340 Display technologies, 190–194 Distinct complex poles, 650–651 Distinct real poles, 645–646 Distributed elements, 358 Domain transformation, 633 Dot convention, 603 Double Data Rate 4 RAM (DDR4RAM), 205

“book” — 2015/5/4 — 7:34 — page 770 — #4

770 Down-conversion, 549 Downswing time constant, 434 Drain, 219 Drift, 21 Drift velocity, 21 Duration, 22 Duration of the pulse, 254 Dynamic circuit, 249 Dynamic RAMs (DRAMs), 204 Dynamic range, 195

E Early time response, 249 Earth ground, 26 Effective value, 462 Electrical engineering for audiophiles, 544–546 Electrical noise, 609 Electrical permittivity, 258 Electrical properties of capacitors, 259–263 of inductors, 270–273 of sea ice measurement, 126–127 Electrical safety, 32–33 Electrical susceptibility, 258 Electric field, 30–31, 258 Electrochemical capacitor, 267 Electrolytic capacitor, 267 Electromagnetic compatibility, 386 Electromagnetic energy, 465 Electromagnetic resonance touchscreens, 395 Electromagnetic spectrum, 465–466 Electron drift, 21 Electron gun, 190 Electronic design automation (EDA), 225 Electronic ink, 192–194 Emitter, 158 Encode, 696 Energy, 577 Energy considerations for magnetically coupled circuits, 615–617 Energy consumption, 34–35 Energy density, 265 Energy dissipated, 525 Energy harvesting, 577 Energy scavenging, 577 Equivalent capacitance, 308

INDEX -equivalent circuit, 614 Equivalent circuits, 35, 67–69, 73–80, 410–414, 613–615, 618–619 resistors and sources in parallel, 73–77 resistors in series, 68–69 sources in series, 69, 73 source transformation, 77–80, 410 Th´evenin equivalent circuit, 410–413 Equivalent-circuit op-amp model, 187–189 Equivalent voltage source, 36 Equivalent resistance Rs , 36 Equivalent resistor, 68 Etching, 136 Euler’s identity, 389 Even function, 683 Even symmetry, 683 Everlasting, 710–711 Expansion coefficient, 645 Exponential form, 698 Exponential Fourier series, 697–699 Exponential function, 256 Exponential waveform, 256–257 Exposure, 137 Extraordinary node, 16–17

F Fall time, 663 Feature size, 294 Feedback, 195 Feedback control, 511 Feedback loop, 70 Feedback resistance, 198 Feedforward control, 511 Ferrite-core inductor, 270 Ferroelectric RAM (FeRAM), 205 Fibrillation, 32 Filter, 501 Filter order, 530–532, 536 Final condition, 280 Final value, 250 First-order circuit, 249, 278, 331 First-order lowpass RC filter, 530–532 First-order RC circuit, 275 Fluid gauge, 302 Forced response, 287 Forcing function, 277

“book” — 2015/5/11 — 17:43 — page 771 — #5

INDEX Forward bias, 59–60, 88, 433 Forward voltage VF , 60 Fourier analysis technique, 674–726 Fourier circuit applications, 690–693 Fourier coefficient, 677 Fourier series analysis technique, 675–677 Fourier series representation, 677–687, 690 amplitude/phase representation, 681–683 odd-function Fourier coefficients, 684–690 sine/cosine representation, 678–681 symmetry considerations, 683–684 Fourier theorem, 675 Fourier transform, 697–701, 704 convergence of Fourier integral, 704 exponential Fourier series, 697–699 nonperiodic waveforms, 699–701 Frequency, 386 Frequency conversion, 549–550 Frequency domain, 396 Frequency domain technique, 386 Frequency filters, 500 Frequency modulation (FM), 548 Frequency response measurement, myDAQ, 757–760 Frequency response of circuits and filters, 500–565 Frequency scaling, 508 Frequency-selective circuits, 501 Frequency shift, 640 Frequency-shift property, 705 Frequency spectrum, 701 Full-wave rectifier, 434 Functional blocks, 227 Functional forms, 516–518 Function generator (FGEN), myDAQ, 753–755 Fundamental angular frequency, 677 Fundamental dimension, 9 Fundamental SI unit, 9 Fuse, 86–87

G Gain-control resistance, 215 Gain factor, 503 Gain roll-off rate, 518 Gate, 219 Genes, 696 Giant magnetoresistance (GMR), 294 Gibbs phenomenon, 687

771 Glasses-free 3-D, 638 Google Earth, 648–649 Gradient coils, 609 Grapher, 96–97 Graphical user interface (GUI), 225 Ground, 26–27 Ground hatch, 301

H Half-power frequency Half-wave rectifier, 434 Hard disk drive (HDD), 293–294 Hardware description languages (HDL), 226 Harmonic, 676 Heterodyne receiver, 548 High-definition television (HDTV), 192 High-frequency gain, 503 High input resistance, 188 Highpass filter, 502, 528–529 High-temperature superconductors, 58 Historical timeline, 4–9 Homogeneity property, 116 Homogeneous, 341 Homogeneous solution, 341 Humidity sensor, 302

I Ideal diode, 88 Ideal current source, 37 Ideal voltage source, 35–36 Idealized response, 503 Ideal op-amp current constraint differentiator, 297 integrator, 295–297 model, 196–198 voltage constraint, 197 Ideal resistor, 54–55 Ideal transformers, 432–433, 617–619 equivalent circuits, 618–619 input impedance, 618 IF, 548 Iff, 390 IF filter/amplifier, 549 Image intensifier, 479–480 Imaginary, 389

“book” — 2015/5/4 — 7:34 — page 772 — #6

772 Impedance in the s-domain, 653–654 Impedance-matching network, 482 Impedance of circuit elements, 397–400 Impedance transformations, 403–410 parallel impedances, 404–407 series impedances, 404–407 Y- transformation, 407–410 Impedance Z, 398 Impedance Zs , 140 Implantation, 136 Impulse function, 631 Indefinite integrals, 736–737 Independent current source, 37–38 Independent voltage source, 35–36 Inductance, 269 Inductive impedance, 403 Inductors, 269–275, 601 electrical properties of inductors, 270–273 in the s-domain, 653 series and parallel combinations of inductors, 273–275 Infrared rays, 466 Infrared touchscreens, 395 Initial condition, 260 Initial time step, 369 Initial value, 250 In parallel, 18 In-parallel connections, 55–56 In phase, 467 Input and output resistances, 141–143 Input impedance, 612–613, 618 Input resistance, 142, 198 In series, 27 In-series connections, 55–56 Instantaneous current, 460 Instantaneous power, 460 Instantaneous values of periodic waveforms, 460–462 Instantaneous voltage, 460 Instrumentation amplifier, 214–216 Integrated circuit, 136 Integrated circuit fabrication process, 136–139 Interconnects, 139 Intermediate frequency, 548 International system of units, 9 Inverse Laplace transform, 635–636 Inverter, 160 Inverting adder, 200 Inverting amplifiers, 198–200

INDEX Inverting input, 185, 188 Inverting pins, 185 Inverting summing amplifier, 200–202, 206 Invoke initial and final conditions, 332 Ionized, 21 Iron-core inductor, 586 Iron-core solenoid, 270 IV analyzer, 234 i–υ relationship, 35, 54

J Junction, 88, 139

K Kilowatt-hours, 29 Kirchhoff’s current law (KCL), 60–62 Kirchhoff’s voltage law (KVL), 62–63 Knee voltage, 88

L Laplace transform, 631 Laplace transform pair, 634 Laplace transform properties, 639–641 frequency shift, 640 time differentiation, 640 time integration, 640–641 time scaling, 639 time shift, 639 Laplace transform technique, 633–636 Larmor frequency, 608 Lattice, 57 Law of conservation of charge, 60 Law of conservation of energy, 62 Law of conservation of power, 29 Least significant bit, 216 Lenticular-lens arrays, 638 Lenz’s law, 609 Light-emitting diodes (LEDs), 59–60, 88, 93–95, 192 Linear and nonlinear elements, 116–117 Linear circuits, 35, 116–117 advantages, 117 homogeneity property, 116 linear and nonlinear elements, 116–117 superposition principle, 116 Linear circuits and source superposition, 133–135, 140

“book” — 2015/5/4 — 7:34 — page 773 — #7

INDEX Linear dynamic range, 195 Linear i–υ relationship, 35 Linear property, 704 Linear region, 54 Linear resistor, 54 Linear transformer, 611 Line current, 572 Line spectrum, 681 Line-to-line voltage, 572 Line voltage, 572 Liquid crystal displays (LCD), 191–192 Load circuit, 142, 151 Load impedance ZL , 140 Loading, 209 Load resistance, 142 Local oscillator, 549 Loop 17 Loosely coupled, 611 Low output resistance, 188 Lowpass filter, 502, 529 Lumped elements, 358

M Maglev trains, 57 Magnetically coupled circuits, 601–629 energy considerations, 615–617 ideal transformers, 617–619 magnetic coupling, 602–607 three-phase transformers, 619–622 transformers, 611–615 Magnetically coupled voltage, 601 Magnetic coupling, 602–607 Magnetic dipole moment, 608 Magnetic field, 608 Magnetic flux, 602 Magnetic-flux linkage, 269, 602 Magnetic permeability, 269, 602 Magnetic resonance imaging (MRI), 57, 608–610 Magnetoresistive RAM (MRAM), 205 Magnitude, 570 Magnitude scaling, 507 Magnitude scaling factor, 507 Magnitude spectrum, 528 Matched load, 483 Matching network, 459

773 Mathematical formulas, 736–737 definite integrals, 737 indefinite integrals, 736–737 trigonometric relations, 736 MATLAB and MathScript, 738–742 basic computation, 739–740 partial fractions, 740–742 MATLAB or MathScript solution, 732 definite integrals, 737 indefinite integrals, 736–737 trigonometric relations, 736 Matrix solution method, 731 Maximum gain, 195 Maximum power transfer, 151–153, 157–158, 476, 481–482 Mean value, 462 Mechanical harvesting, 578 Mechanical load, 70 Mechanical stress, 70, 90 MEMS, 14 Mesh, 17, 123 Mesh analysis by inspection, 132–133 Mesh-current method, 123–125, 128–129 Metal-oxide semiconductor field-effect transistor (MOSFET), 219 Mica capacitor, 259 Micro- and nanotechnology, 10–14 Microchannel plate (MCP), 480 Microelectromechanical systems (MEMS), 14, 336 Micromechanical sensors and actuators, 336–339 capacitive MEMS actuator, 338–339 capacitive MEMS accelerometer, 336–337 microelectromechanical systems (MEMS), 336 Miniaturized energy harvesting, 577–578 mechanical harvesting, 578 radio frequency scavenging, 578 thermoelectric harvesting, 577–578 Mixed-signal circuits, 155–156, 713 Mixer, 140–141 Modulation, 547–548 Moisture and chemical sensors, 71–72 Moore’s law, 10–13 MOSFET, 219, 223–224, 688 MOSFET gain constant, 220 Most significant bit, 216 Motherboard, 306

“book” — 2015/5/11 — 17:44 — page 774 — #8

774 Multisim, 91–92, 95–99, 161–163, 225–227, 231–234, 437–442, 482–485, 550–554, 662–664, 713–717, 733–735 ac analysis, 437–442 circuit response, 310–312, 369–373 circuit simulation software, 225–227 dependent sources, 97–99 digital inverter, 231–234 drawing the circuit, 91–92, 95–96 sigma-delta modulator, 713–717 nodal analysis with, 161–163 nontrivial inputs, 662–664 op amps and virtual instruments, 230–231 overview, 733–735 power measurement, 482–485 solving the circuit, 96–97 spectral response, 550–554 Mutual capacitive sensing, 393 Mutual inductance, 269, 602 Mutual inductance voltage, 602 Mutual voltage, 603 myDAQ quick reference guide, 743–760 arbitrary waveform generator, 756–757 current measurement, 747–750 frequency response measurement, 757–760 function generator (FGEN), 753–755 NI myDAQ as current source, 751–753 protoboard/breadboard, 750–751 resistance measurement, 746–747 time-varying voltage measurement, 755–756

N Nanocapacitor, 259 Nano RAM (RAM), 205 National Institute for Standards and Technology (NIST), 424 Natural response, 276, 287, 341 Natural response, RL circuit, 287–288 Near-IR, 478 Negative feedback, 195–196 Negative phase sequence, 570 Negative saturation, 186 Nepers/second (Np/s), 342 Net charge, 20, 24 Network, 573

INDEX Neural interface, 229 Neural probes, 229 Neural stimulation and recording, 363–365 Neurons, 229 Neutral node, 572 Neutral terminal, 570 Neutral wire, 570 NI myDAQ as current source, 751–753 nmax -truncated series, 680 NMOS versus PMOS transistors, 221–223 Nodal analysis by inspection, 130–132 with Multisim, 161–163 Node, 16 Node-voltage method, 117–123 circuits with dependent sources, 120–121 general procedure, 117–120 supernodes, 121–123 Node voltages, 27 Noise-cancellation headphones, 509–511 Noninverting amplifier, 188, 197 Noninverting summer, 201–202, 206 Nonlinear elements, 116–117 Nonperiodic waveforms, 250–258, 699–701 exponential waveform, 256–257 and Fourier analysis technique, 699–701 pulse waveform, 253–256 ramp-function waveform, 252–253 step-function waveform, 250–252 Nonplanar, 19 Normalized power, 512 Norton’s theorem, 149 NOT logic gate, 156 npn configuration, 158 n-type semiconductor, 88 Nuclear magnetic resonance (NMR), 608 Null, 701

O Odd function, 683 Odd-function Fourier coefficients, 684–687, 690 Odd symmetry, 683 Offset voltage VF , 88

“book” — 2015/5/11 — 17:44 — page 775 — #9

INDEX Ohm’s law, 51–56, 59–60 conductance, 69 ideal resistor, 54–55 in-parallel connections, 55–56 in-series connections, 55–56 light-emitting diodes (LEDs), 59–69 resistance, 52–54 superconductivity, 57–58 One-sided excitation, 711 One-sided transform, 634 One-way valve, 88 Op-amp integrator, 295 Open-loop gain, 186 Operational amplifiers (op amps), 16, 38, 183–247 characteristics, 184–189 computer memory circuits, 203–205 current constraint, 197 difference amplifier, 206–207 digital-to-analog converters (DAC), 216–219 ideal model, 196–198 instrumentation amplifier, 214–216 inverting amplifiers, 198–200 inverting summing amplifier, 200–202, 206 Multisim analysis, 230–234 negative feedback, 195–196 signal-processing circuits, 209–214 voltage follower/buffer, 208–209 Ordinary node, 16 Organic LEDs (OLEDs), 192 OR logic gate, 156, 540 Oscillation frequency, 386 Oscillator, 423 Oscilloscope, 230 Output resistances, 141–143 Overcurrent, 86 Overdamped response, 340–346 Overloading, 208 Oxidation, 139 Oxide layer, 139

P Parallax barrier 3-D, 638 Parallel-plate capacitor, 258 Parallel RLC circuit, 353–355, 359 Parasitic capacitance, 305–310 Paresthesia, 364

775 Parseval’s theorem, 710 Partial fraction expansion, 644–647, 650–652 distinct complex poles, 650–651 distinct real poles, 645–646 repeated complex poles, 651–652 repeated real poles, 646–647, 650 Particular solution, 341 Passband, 502 Passive filters, 522–530 bandpass filter, 523–528 bandreject filter, 529–530 highpass filter, 528–529 lowpass filter, 529 Passive RFID, 357 Passive sign convention, 29 Path, 17 PCB layout, 15 Peak-to-peak ripple voltage, 436 Peak value, 386 Percent clipping, 245 Percolation threshold, 126 Perfectly coupled, 611 Periodic excitation, 674 Periodic function, 675 Periodicity property, 460 Periodic waveforms, 249, 460–463 average values, 460–462 instantaneous values, 460–462 root-mean-square value, 462–463 Period (of a cycle), 386 Permeability of free space, 611 Permittivity capacitive sensors applications, 301–303 Perpendicular magnetic recording (PMR), 294 Phase, 388 Phase angle, 387 Phase current, 72 Phase lag, 387 Phase lead, 387 Phase representation, 681 Phase-shift circuits, 416–420 Phase-shift oscillator, 456 Phase spectrum, 681 Phase transition, 57 Phase voltage, 570, 572 Phasor counterpart, 396 Phasor diagrams, 413–416 Phasor domain, 396–403

“book” — 2015/5/4 — 7:34 — page 776 — #10

776 Phasor-domain techniques, 420–422, 425–429 Phasor versus Laplace versus Fourier, 710–711 Photon, 465 Piezoresistive coefficient, 70, 90 Piezoresistive effect, 70 Piezoresistor, 54 Planar circuits, 19–20 Planck’s constant, 465 Plasma displays, 192 Plastic-foil capacitor, 259 PMOS transistors, 221–223 pn-junction diode, 88 pnp configuration, 158 Polar form, 389 Polarization, 638 Polarizing 3-D, 638 Pole @ origin factor, 516–517 Pole factor, 530 Poles, 515, 645 Positive feedback, 195 Positive (123) phase sequence, 569 Positive saturation, 186 Potential difference, 25 Potentiometer, 54 Power, 28–29, 34–35, 265 Power factor, 472–476, 588 compensation, 474–476 significance, 473–474 Power factor angle, 467 Power factor compensation, 588–591 Power generation station, 586–587 Power grid, 567 Power in balanced three-phase networks, 582–588 -load configuration, 583 total instantaneous power, 583–585, 588 Y-load configurations, 582–583 Power measurement in three-phase circuits, 591–594 Power rating, 55 Power supply circuits, 432–437 ideal transformers, 432–433 rectifiers, 433–434 smoothing filters, 434–436 voltage regulator, 436–437 Prefixes, 9 Pressure touchscreens, 395 Primary port, 602 Primary winding, 432

INDEX Printed circuit board, 15, 225 Printed conducting lines, 15 Proper rational function, 646 Protoboard/breadboard, myDAQ, 750–751 Prototype model, 507 p-type semiconductor, 88 Pulsed electromagnetic field (PEMF), 364 Pulse repetition frequency, 305 Pulse waveform, 253–256

Q Quadratic-pole factor, 650 Quadratic-zero factor, 518 Quality factor, 524 Quantization error, 155 Quartz, 423 Quartz crystals and piezoelectricity, 423 Quasi-supernode, 121

R Radio-frequency (RF), 548, 608 Radio frequency identification (RFID), 331, 356, 370–373, 386, 578 Radio frequency scavenging, 578 Radio waves, 466 Ramp-function waveform, 252–253 Random-access memories (RAMs), 204–205 RC and RL first-order circuits, 248–329 RC circuit, 250, 275 RC circuit response, 275–287 RCL circuits, 275, 330–384 RC op-amp circuits, 295–300, 304–305 ideal op-amp differentiator, 297 ideal op-amp integrator, 295–297 other op-amp circuits, 297–300, 304–305 Reactance, 403 Reactive power, 468 Realistic current source, 38 Realizable circuits, 76 Real power, 582 Real voltage source, 36 Receiver, 548

“book” — 2015/5/4 — 7:34 — page 777 — #11

INDEX Rectangular form, 389 Rectangular function, 254 Rectangular pulse, 253 Rectenna, 578 Rectifiers, 433–434 Reflected impedance, 614 Regenerative receiver, 548 Region of convergence, 635 Relationship between u(t) and δ(t), 632 Relative permittivity, 258 Relative phasor diagram, 414 Relative power, 512 Repeated complex poles, 651–652 Repeated real poles, 646–647, 650–652 Residue method, 645 Resistance, 52–54 Resistance matrix, 132 Resistance measurement, myDAQ, 746–747 Resistive circuits, 50–114 Resistive sensors, 70–72 moisture and chemical, 71–72 thermistor, 70–71 Resistive touchscreens, 393 Resistivity, 51 Resistor in the s-domain, 653 Resistors and sources in parallel, 73–77 Resistors in series, 68–69 Resonance condition, 523 Resonant frequency, 342, 523 Resonator and clock advances, 424 Reversal property, 710 Reverse bias, 59, 88 RF, see Radio frequency RFID, see Radio frequency identification (RFID) RFID tags and antenna design, 356–358 antennas, 357–358 applications, 356–357 operation, 357 Rheostat, 54 Ripple, 436 Rise time, 227, 663 RLC bandpass filter, 536 RL circuit, 250, 275 RL circuit response, 287–292, 295 rms value, 462–463 Room-temperature superconductors, 58 Rotor, 568

777 Round-off error, 154 RTh —equivalent resistance method, 146–147 RTh —external source method, 147–149 R–2R ladder, 216 Rule of fives, 126

S Sampling property, 632, 633 Scaled inverting adder, 200 Scale of things, 10 Scaling factor, 296 Scaling trends and nanotechnology, 13–14 Schematic capture window, 91 Schmidt triggers, 717 s-domain circuit analysis, 655–662 s-domain circuit element models, 652–654 capacitor in the s-domain, 653 impedance in the s-domain, 653–654 inductor in the s-domain, 653 resistor in the s-domain, 653 Secondary port, 602 Secondary winding, 432 Second-order circuit, 331 Second-order lowpass filter, 532 Self-capacitive sensing, 393 Self-inductance, 269, 602 Semiconductor memories, 203 Semiconductors, 51 Sensor, 15–16 Sensor pad, 301 Sensory and motor prostheses, 364 Series and parallel combinations of capacitors, 263–264, 268–269 of inductors, 273–275 of resistors, 68–69, 73–77 Series impedances, 404–407 Series RLC circuit, 334–335, 340 Shannon-Hartley theorem, 689 Shift properties, 705 Shingled magnetic recording (SMR), 294 Short circuit, 28 Siemen, 60 Sigma-Delta Modulator, 713–717 Signal, 70 Signal and noise in communication, 688–689 Signal-processing circuits, 209–214

“book” — 2015/5/4 — 7:34 — page 778 — #12

778 Signal-to-noise ratio, 688–689 Signum function, 706 Simple-pole factor, 519 Simple-zero factor, 517–518 Sinc function, 699 Sine/cosine representation, 678–681 Single-phase equivalent circuits, 579–582 Single-phase generators, 572 Single-pole double-throw (SPDT), 28 Single-pole highpass filter, 537–538 Single-pole lowpass filter, 537 Single-pole single-throw (SPST), 28 Singularity function, 635 Sinusoidal signals, 386–389 Smoothing filters, 434–436 Software radio, 550 Solenoid, 269 Solid state, 203 Source, 15–18 Source circuit, 151 Source-free, 276 Source-free, first-order differential equation, 277 Source impedance Zs , 140 Sources in series, 69, 73 Source-load configurations, 572–573 balanced conditions, 572–573 balanced networks, 576, 579–582 Y and  notation, 572 Y-Y configurations, 574–576 Source superposition, 133–135, 140 Source transformation, 77–80, 410 Source-transformation principle, 410 Source vector, 131 Sources in series, 69, 73 Spacing between adjacent harmonics, 699 Spatial filters, 535 Speakers, 545–546 Spectral distortion, 544 Spectral filters, 533–535 SPICE, 91 Spinal cord stimulator, 364 Spring constant k, 337 Square wave, 89, 110 Standard form, 515 Static RAMs, 204 Stator, 569 Stator coils, 586

INDEX Steady-state, 233 Steady-state component, 249, 711 Steady-state response, 249–250 Steady-state solution, 341 Step-down transformer, 567, 617 Step function, 250, 631 Step-function waveform, 250–252 Step response, general form RC circuit, 279–287 RL circuit, 288–292, 295 Step-up transformer, 567, 617 Stereopsis, 637 Stimulus, 70 Stopband, 502 Substrate, 136 Summing amplifier, 200–201 Supercapacitors, 259, 265–267 Superconducting electromagnet, 608 Superconducting Quantum Interference Devices (SQUIDs), 57 Superconductivity, 57–58 Superconductor, 51 Superheterodyne receiver, 548–549 Supermeshes, 128–129 Supernodes, 121–123 Superposition principle, 116 Susceptance, 403 Switches, 28, 39–40, 310–311 Switching frequency (speed), 305 Symbols, quantities, and units, 727–728 Symmetry considerations, 683–684 Synchronization, 549 Synchronous Graphics RAM (SGRAM), 205 Synthesis, 3–4 Synthetic biology, 695–696 System, 2

T Technology circuit fabrication process, 136–139 T-equivalent circuit, 613 Thermal energy, 478 Thermal-infrared imaging, 477–479 Thermal-IR, 478 Thermistor, 54 Thermistor sensors, 70–71 Thermoelectric harvesting, 577–578

“book” — 2015/5/4 — 7:34 — page 779 — #13

INDEX Thermoelectric materials, 577 Th´evenin equivalent circuit, 145, 150, 157, 218, 283, 410–413 Th´evenin impedance, 482 Th´evenin’s theorem, 143–144 Th´evenin voltage, 149 Thin-film head, 294 Thin-film transistor, 191 3-D mapping 3-D modeling tools, 227 3-D TV, 637–638 Three-phase ac generator, 568 Three-phase circuits, 566–600 Three-phase network, 566 Three-phase power, 566 Three-phase transformers, 619–622 Three-wire single phase, 568 Tightly coupled, 611 Time constant, 256, 289 Time-dependent sources, modeling, 311–312 Time differentiation, 640 Time-domain/phasor-domain correspondence, 396–397 Time period, 348 Time scaling, 639 Time shift, 388, 639 Time-shifted ramp function, 252 Time-shifted step function, 251 Time-shift property, 639 Time-varying function, 396 Time-varying voltage measurement, myDAQ, 755–756 Total response, 674 Touch controller IC chips, 395 Touchscreen, 393 Touchscreens and active digitizers, 393–395 Transducer, 303, 336 Transduction, 336, 577 Transfer function, 501–507 Transformation between balanced loads, 576 Transformation between balanced sources, 576 Transformers, 567, 611–615 coupling coefficient, 611–612 equivalent circuits, 613–615 input impedance, 612–613 Transient component, 711 Transient response, 249, 341 Transistor, 4 Transmission line, 306, 566, 574

779 Transmission velocity, 21 Transmission window, 688 Trigonometric relations, 736 Trivial resonance, 504 True phase angle, 416 Truncated series, 680 Tuned-radio frequency, 548 Tuner, 548 Tunneling magnetoresistance (TMR), 294 Turns ratio, 617 Tweeters, 546 Two parallel lines, 617 Two-sided transform, 634 Two-source circuit, 65 Two-wattmeter method, 592

U Ultracapacitor, 266 Ultraviolet imaging, 480 Ultraviolet rays, 465–466 Uniqueness property, 634 Unit, 9 Unit impulse function, 631–633 Unit rectangular function, 254 Units, dimensions, and notation, 9, 15 Unit step function, 313 Unity gain, 195 Unity gain amplifier, 208 Unity input, 501 Universal property, 639 Unrealizable circuit, 73 Up-conversion, 549 Upswing time constant, 434

V VAR, see Volt-ampere reactive Variable resistance, 54 Very large scale integrated circuits (VLSI), 225 Virtual globe, 648 Visible light rays, 466 Voltage, 25–28 Voltage coil, 591 Voltage constraint, 197 Voltage-controlled voltage source (VCVS), 37, 39 Voltage difference, 27 Voltage divider, 69, 100

“book” — 2015/5/4 — 7:34 — page 780 — #14

780 Voltage drop, 25 Voltage follower/buffer, 208–209 Voltage rails, 187 Voltage regulator, 436–437 Voltage rise, 25, 29 Voltage transfer function, 501 Voltage transformers, 601, 602 Voltage vector action potential, 131 Volt-ampere, 468 Volt-ampere reactive, 468 Voltmeter, 143

INDEX

X XOR logic gate, 156 X-rays, 465

Y Y configuration, 586 Y- transformation, 407–410. See also Wye–Delta (Y–) transformation Y-load configurations, 582–583 Y-Y configurations, 574–576

W

Z

Wafer, 136 Wattmeter, 591 Wave-particle duality, 465 Wheatstone bridge, 84–86 Windings, 432 Woofers, 546 Word, 154 Wye–Delta (Y–) transformation, 80–84

Zener diode, 436 Zener-diode resistance, 436 Zener voltage, 436 Zero @ origin factor, 516, 517 Zeroes, 515 Zero factor, 517–519

CIRCUIT ANALYSIS AND DESIGN Fawwaz T. Ulaby, Michel M. Maharbiz, & Cynthia M. Furse

Fawwaz Ulaby is the Emmett Leith Distinguished Professor of Electrical Engineering and Computer Science and former Vice President of the University of Michigan. He is a member of the National Academy of Engineering and recipient of the IEEE James H. Mulligan, Jr. Education Medal. His Applied Electromagnetics textbook is used at over 100 US universities. Michel M. Maharbiz is Professor of Electrical Engineering and Computer Science at the University of California, Berkeley. Dr. Maharbiz has been a GE Scholar, an Intel IMAP Fellow and is currently a Bakar Fellow. He was awarded National Instrument’s Excellence in Engineering Education Award in 2013. His research interests include building micro/nano interfaces to cells and organisms and exploring bio-derived fabrication methods. Cynthia M. Furse is Professor of Electrical and Computer Engineering at the University of Utah. She has received numerous teaching awards, and is a Fellow of the IEEE and the National Academy of Inventors. Her research is in bioelectromagnetics (how electromagnetic fields interact with the body).